Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-06-10T01:17:12.292Z Has data issue: false hasContentIssue false

8 - Soil Moisture

Published online by Cambridge University Press:  08 February 2019

Gordon Bonan
Affiliation:
National Center for Atmospheric Research, Boulder, Colorado

Summary

Water flows from high to low potential as described by Darcy’s law. The Richards equation combines Darcy’s law with principles of water conservation to calculate water movement in soil. Particular variants of the Richards equation are the mixed-form, head-based, and moisture-based equations. Water movement is determined by hydraulic conductivity and matric potential, both of which vary with soil moisture and additionally depend on soil texture. This chapter reviews soil moisture and the Richards equation. Numerical solutions are given for the various forms of the equation.
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019

Chapter Overview

Water flows from high to low potential as described by Darcy’s law. The Richards equation combines Darcy’s law with principles of water conservation to calculate water movement in soil. Particular variants of the Richards equation are the mixed-form, head-based, and moisture-based equations. Water movement is determined by hydraulic conductivity and matric potential, both of which vary with soil moisture and additionally depend on soil texture. This chapter reviews soil moisture and the Richards equation. Numerical solutions are given for the various forms of the equation.

8.1 Introduction

The region of soil between the ground surface and the water table is known as the unsaturated, or vadose, zone, and the water held in this zone is called soil moisture (Figure 8.1). The water content of the vadose zone is dynamic, ranging from saturation in the upper soil layers near the surface during infiltration to nearly dry in prolonged absence of rainfall as plant roots extract water during transpiration. The vertical profile of soil water is a particularly important determinant of land–atmosphere coupling. A dry surface layer develops in the absence of rainfall, and this dry layer impedes soil evaporation. Conversely, plant roots can extend deep in the soil to sustain transpiration during dry periods. Below the vadose zone lies saturated groundwater, and soil moisture also controls the fluxes of water between the vadose zone and groundwater.

Figure 8.1 Water flows in a soil column extending from the ground surface to the water table.

The first models of the land surface used in climate simulations ignored the complexity of the hydrologic cycle and instead abstracted it using the bucket analogy in which soil is treated as a bucket that fills from precipitation, empties from evapotranspiration, or spills over as runoff as described in Chapter 7. In fact, however, storage of water and its movement in soil is much more complex. When soil is wet, water is loosely held in soil and quickly drains due to the force of gravity. When soil is dry, water movement becomes more difficult, and at some critical amount the water is strongly bound to soil particles and can no longer be removed. Figure 8.2 illustrates the dynamics of water movement during infiltration into initially dry soil. A distinct wetting front moves progressively downward over time. The upper soil becomes saturated with water while the deeper soil remains dry. In the sandy soil shown in Figure 8.2, the upper 50–60 cm become saturated after 42 minutes (0.7 h). This dynamics is explained from physical principles using Darcy’s law and the Richards equation.

Figure 8.2 Soil water movement during infiltration into sand. Shown are the initial moisture profile (θ = 0.1θ=0.1) and profiles in increments of 0.1 h.

8.2 Measures of Soil Moisture

A typical soil consists of solid particles of varying size and shape that are interconnected by pores. Water completely fills these pores when the soil is saturated, or the pores consist mainly of air when the soil is dry. Most conditions in the field are in-between, and soil is a mix of solid particles, water, and air (Figure 8.3). The bulk density of a soil ρbρb (kg m–3) is the mass of soil solids per volume of soil so that

ρb=mass of solidsvolume of soil=msV.(8.1)

The bulk volume of soil consists of the total volume of solids and pore space (V = Vs + VpV=Vs+Vp). The particle density ρsρs (kg m–3) is the mass of soil solids per volume of soil solids whereby

ρs=mass of solidsvolume of solids=msVs.(8.2)

If a volume of soil 10 cm × 10 cm × 10 cm has a dry mass of 1.325 kg, its bulk density is 1325 kg m–3. If the pore space comprises one-half of this volume, Vs = 0.5VVs=0.5V, and the particle density is 2650 kg m–3. A typical particle density is, in fact, 2650 kg m–3. The fraction of the soil volume comprising pores, known as porosity, is (V − Vs)/VVVs/V. When saturated, water fills all the pores so that porosity is also the volumetric water content at saturation θsatθsat. Porosity is calculated from bulk density and particle density by

θsat=1ρbρs.(8.3)

Figure 8.3 Depiction of soil with total volume VV comprising soil solids VsVs, water VwVw, and air VaVa. The total volume of pore space is Vp = Vw + VaVp=Vw+Va. The total soil mass consists of soil solids with mass msms and water with mass mwmw.

The amount of water can be measured by volume (m3 H2O) and equivalently by mass per area (kg H2O m–2) or depth (m H2O). These are related by the density of water (ρwat = 1000ρwat=1000kg m–3) so that mass per area = depth × densitymassperarea=depth×density. One kilogram of water spread over an area of one square meter (1 kg m–2) is equivalent to a depth of 1 mm and a volume of 0.001 m3. A common measure of soil moisture is volumetric water content (m3 H2O m–3 soil). Volumetric water content is

θ=volume of watervolume of soil=VwV.(8.4)

Volumetric water content is also the depth of water per unit depth of soil. A soil with thickness ΔzΔz m contains θ ΔzθΔz m of water. The mass of water per area WW (kg m–2) in a volume of soil with volumetric water content θθ is

W = θ Δwat
W=θΔzρwat.
(8.5)

Another measure of soil moisture is gravimetric (mass) water content (kg H2O kg–1 dry soil). Gravimetric water content is

θm=mass of watermass ofdrysoil=mwms.(8.6)

Volumetric water content is related to gravimetric water content by the density of water and the bulk density of soil as

θ=θmρbρwat.(8.7)

Another measure is the effective saturation, which is defined as the soil moisture θθ above some residual amount θresθres relative that at saturation:

Se=θθresθsatθres.(8.8)

Table 8.1 provides example calculations for these various measures of soil moisture.

Table 8.1 Various measures of soil water for a volume of soil with dimensions 10 cm × 10 cm × 10 cm that has a mass of 1.7 kg when wet, 1.45 kg when dry, and particle density ρs = 2650ρs=2650 kg m–3

QuantityAmount
Mass of watermw = 1.7 kg − 1.45 kg = 0.25mw=1.7kg1.45kg=0.25kg
Mass of soil solidsms = 1.45ms=1.45kg
Bulk volume of soilV = 0.1 m × 0.1 m × 0.1 m = 0.001V=0.1m×0.1m×0.1m=0.001m3
Bulk densityρb = ms/V = 1450ρb=ms/V=1450 kg m–3
Porosityθsat = 1 − ρb/ρs = 0.453θsat=1ρb/ρs=0.453
Volume of waterVw = mw/ρwat = 0.00025Vw=mw/ρwat=0.00025 m3
Gravimetric water contentθm = mw/ms = 0.172θm=mw/ms=0.172 kg kg–1
Volumetric water contentθ = Vw/V = θmρb/ρwat = 0.25θ=Vw/V=θmρb/ρwat=0.25 m3 m–3
Water content relative to saturationθ/θsat = 0.552θ/θsat=0.552
Depth of waterθ Δz = 0.25 × 0.1 m = 0.025θΔz=0.25×0.1m=0.025 m
Mass of water per areaW=0.25kg0.1m×0.1m=θΔzρwat=25 kg m–2

8.3 Matric Potential and Hydraulic Conductivity

Water is tightly held by the surfaces of soil particles. This creates a negative pressure, or suction, called matric potential ψψ that binds water to the soil. (The symbol hh is commonly used in soil science and hydrology to denote this as pressure head.) The matric force is positive when represented as suction and negative when given as potential. Matric potential varies with soil moisture. Relatively weak suction is exerted on water when soil is saturated (matric potential is high), but suction increases sharply (matric potential decreases) as soil becomes drier and strong forces bind water in small pores. This relationship between ψψ and θθ is quite nonlinear and varies depending on soil texture (Figure 8.4). Water is loosely held in sandy soils (low suction) and tightly held in clay soils (high suction).

Figure 8.4 Water content in relation to matric potential given here as suction (ψψ).

Shown is the van Genuchten (Reference van Genuchten1980) θ(ψ)θψrelationship for Berino loamy fine sand (θres = 0.0286θres=0.0286, θsat = 0.3658θsat=0.3658, α = 0.028α=0.028 cm–1, n = 2.239n=2.239) and Glendale silty clay loam (θres = 0.106θres=0.106, θsat = 0.4686θsat=0.4686, α = 0.0104α=0.0104 cm–1, n = 1.3954n=1.3954) from Hills et al. (Reference Hills, Hudson, Porro and Wierenga1989).

The dependence between ψψ and θθ is referred to as the soil moisture retention curve and is described mathematically by equations that relate θθ to ψψ or, equivalently, ψψ to θθ, denoted θ(ψ)θψ or ψ(θ)ψθ, respectively. Table 8.2 gives three common relationships. Brooks and Corey (Reference Brooks and Corey1964, Reference Brooks and Corey1966) related ψψ to the effective saturation SeSe. In this equation, ψbψb and cc are empirical parameters used to fit the data; ψbψb is the air entry water potential and is the value of ψψ at which Se = 1Se=1; cc is referred to as the pore-size distribution index. Campbell (Reference Campbell1974) proposed a similar relationship but with θres = 0θres=0, in which case ψbψb can be thought of as the matric potential at saturation. Van Genuchten (Reference van Genuchten1980) developed another widely used soil moisture retention curve. In this relationship, the empirical parameter αα is the inverse of the air entry potential (α = 1/ ∣ ψbα=1/ψb), and nn is the pore-size distribution index.

Table 8.2 Soil moisture retention and hydraulic conductivity functions

θ(ψ)θψK(θ)Kθ
(a) Brooks and Corey (Reference Brooks and Corey1964, Reference Brooks and Corey1966)
Se=θθresθsatθres=ψψbcK=KsatSe2/c+3
=cθsatθresψbψψbc1dK=Ksatθsatθres2/c+3Se2/c+2
(b) Campbell (Reference Campbell1974)
θθsat=ψψsat1/bK=Ksatθθsat2b+3
=θsatbψsatψψsat1/b1dK=Ksat2b+3θsatθθsat2b+2
(c) van Genuchten (Reference van Genuchten1980)
Se=θθresθsatθres=1+αψnm, m = 1 − 1/nm=11/nK=KsatSe1/211Se1/mm2
=αmnθsatθresαψn11+αψnm+1dK=Ksatθsatθresf22Se1/2+2Se1/m1/2f1Se1/m1m, f=11Se1/mm

Note: (a) Se = 1Se=1 and K = KsatK=Ksat for ψ>ψbψ>ψb. (b) θ/θsat = 1θ/θsat=1 and K = KsatK=Ksat for ψ>ψsatψ>ψsat. (c) Se = 1Se=1 and K = KsatK=Ksat for ψ>0ψ>0.

Parameter values vary depending on soil texture, and various so-called pedotransfer functions relate hydraulic parameters to discrete texture classes or as continuous functions of sand, clay, or other soil properties. The Brooks and Corey (Reference Brooks and Corey1964, Reference Brooks and Corey1966) parameters, for example, can be related to sand, clay, and porosity (Rawls and Brakensiek Reference Rawls, Brakensiek, Jones and Ward1985; Rawls et al. Reference Rawls, Ahuja, Brakensiek, Shirmohammadi and Maidment1993). Clapp and Hornberger (Reference Clapp and Hornberger1978) estimated parameters for various soil texture classes using the Campbell (Reference Campbell1974) relationship, and Cosby et al. (Reference Cosby, Hornberger, Clapp and Ginn1984) subsequently related θsatθsat, ψsatψsat, and bb to the sand and clay content of soil. The van Genuchten (Reference van Genuchten1980) parameters can be difficult to estimate (Carsel and Parrish Reference Carsel and Parrish1988; Leij et al. Reference Leij, Alves, van Genuchten and Williams1996; Schaap et al. Reference Schaap, Leij and van Genuchten1998, Reference Schaap, Leij and van Genuchten2001). Table 8.3 gives representative values for soil texture classes.

Table 8.3 Parameter values for the Campbell (Reference Campbell1974) and van Genuchten (Reference van Genuchten1980) θ(ψ)θψ and K(θ)Kθ functions arranged by soil texture

Soil typeCampbellvan Genuchten
θsatθsatψsatψsat (cm)bbKsatKsat (cm h–1)θsatθsatθresθresαα (cm–1)nnKsatKsat (cm h–1)
Sand0.395–12.14.0563.360.430.0450.1452.6829.70
Loamy sand0.410–9.04.3856.280.410.0570.1242.2814.59
Sandy loam0.435–21.84.9012.480.410.0650.0751.894.42
Silt loam0.485–78.65.302.590.450.0670.0201.410.45
Loam0.451–47.85.392.500.430.0780.0361.561.04
Sandy clay loam0.420–29.97.122.270.390.1000.0591.481.31
Silty clay loam0.477–35.67.750.610.430.0890.0101.230.07
Clay loam0.476–63.08.520.880.410.0950.0191.310.26
Sandy clay0.426–15.310.40.780.380.1000.0271.230.12
Silty clay0.492–49.010.40.370.360.0700.0051.090.02
Clay0.482–40.511.40.460.380.0680.0081.090.20

Note: Soils are arranged from least to most clay.

Source: Campbell (Reference Campbell1974) parameters from Clapp and Hornberger (Reference Clapp and Hornberger1978); van Genuchten (Reference van Genuchten1980) parameters from Carsel and Parrish (Reference Carsel and Parrish1988) and Leij et al. (Reference Leij, Alves, van Genuchten and Williams1996).

Water held in soil is subjected to two forces, or potentials. The force of gravity pulls water downward, denoted as gravitational potential. Gravitational potential is the height zz above some arbitrary reference height. The second force is the matric potential, which holds water to the soil particles. The total potential, also known as hydraulic head, is ψ + zψ+z, and this is the work per unit weight required to move an amount of water at some elevation and matric potential to another position in the soil with a different potential. Soil water potential as used in this chapter has the units of energy per unit weight and has the dimension of length (m):

energyweight=JN=NmN=m

Other units for potential are energy per unit mass (J kg–1) or energy per unit volume (J m–3), and this latter expression is also the units of pressure (Pa). Soil water potential is converted from m to J kg–1 by multiplying by gravitational acceleration gg and to J m–3 by multiplying by ρwatgρwatg:

J kg−1 = 
Jkg1=
J m−3 = Pa = ρwat
Jm3=Pa=ρwat.

A wet soil with a matric potential of –0.01 MPa has a suction of approximately 1000 mm; a dry soil with –1.5 MPa has a suction of approximately 150 m.

Hydraulic conductivity governs the rate of water flow for a unit gradient in potential. Hydraulic conductivity decreases sharply as soil becomes drier because suction increases and because the pore space filled with water becomes smaller and discontinuous. This relationship is nonlinear and varies with soil texture (Figure 8.5). Sandy soil has a higher conductivity than clay soil. The relationship between KK and θθ is not independent of the relationship between θθ and ψψ, and the derivation of K(θ)Kθ requires an expression for θ(ψ)θψ (Figure 8.4). This expression is given in terms of the effective saturation SeSe, or θ/θsatθ/θsat in the Campbell (Reference Campbell1974) equation, so that hydraulic conductivity can be equivalently expressed as K(ψ)Kψ. The expressions for hydraulic conductivity also require KsatKsat, the hydraulic conductivity at saturation. The term Se1/2 in the van Genuchten (Reference van Genuchten1980) equation for hydraulic conductivity is a common form, but the exponent can vary (Schaap et al. Reference Schaap, Leij and van Genuchten2001). Better fit to data can be achieved by replacing KsatKsat with a curve fitting parameter K0K0, but this has a value that is usually less than KsatKsat so that K(θ)≠KsatKθKsat at θsatθsat (Schaap et al. Reference Schaap, Leij and van Genuchten2001).

Figure 8.5 Hydraulic conductivity in relation to water content. Shown is the van Genuchten (Reference van Genuchten1980) K(θ)Kθrelationship for Berino loamy fine sand and Glendale silty clay loam. Parameter values are as in Figure 8.4 and KsatKsat= 22.54 and 0.55 cm h–1, respectively.

Question 8.1 Graph and compare the van Genuchten (Reference van Genuchten1980) and Campbell (Reference Campbell1974) relationships for θ(ψ)θψ and K(θ)Kθ for sandy loam, loam, and clay loam using parameter values in Table 8.3. Describe differences in the shape of these relationships.

8.4 Richards Equation

Darcy’s law describes water flow. In the vertical dimension, the rate of water movement is

Q=Kθψ+zz=Kθψz+1=KθψzKθ.(8.9)

This is a form of Fick’s law and relates the rate of flow to the product of the hydraulic conductivity and the vertical gradient in water potential. The flux QQ is the volume of water (m3) flowing through a unit cross-sectional area (m2) per unit time (s) and has the dimensions length per time (m s–1). Hydraulic conductivity has the same units, and K(θ)Kθ denotes that hydraulic conductivity depends on soil moisture. The total potential ψ + zψ+z has dimensions of length (m). It governs water movement so that water flows from high to low potential. The vertical depth zz is taken as positive in the upward direction so that z = 0z=0 is the ground surface and z < 0z<0 is the elevation relative to the surface with greater depth into the soil. The matric potential has values ψ < 0ψ<0 for unsaturated soil and ψ ≥ 0ψ0 for saturated soil. The negative sign in (8.9) ensures that downward water flow is negative and upward flow is positive.

The flow of water given by Darcy’s law depends strongly on soil moisture. The dominant force causing water to move in a wet soil is the gravitation potential. Water near the surface has a higher gravitational potential than water deeper in the soil, and because water flows from high potential to low potential, it flows downward from the force of gravity. This is given by zz in (8.9), which is the height relative to the ground surface. In drier soils, matric potential decreases, and the adsorptive force binding water to soil particles generally exceeds the gravitational force pulling water downward. This reduces the rate of water flow. The lower hydraulic conductivity in dry soils also restricts water movement.

An equation for the time rate of change in soil moisture is obtained from principles of conservation similar to that for soil temperature. Consider a volume of soil with horizontal area ΔxΔyΔxΔy and thickness ΔzΔz and in which water flows only in the vertical dimension (Figure 8.6). The mass flux of water (kg s–1) entering the soil across the cross-sectional area ΔxΔyΔxΔy is ρwatQinΔxΔyρwatQinΔxΔy, and the flux out of the soil is similarly ρwatQoutΔxΔyρwatQoutΔxΔy. Conservation requires that the difference between the flux of water into and out of the soil equals the rate of change in water storage. The mass of water in the soil volume is ρwatθΔxΔyΔzρwatθΔxΔyΔz so that the change in soil water over the time interval ΔtΔt is

ρwatΔθΔtΔxΔyΔz=ρwatQinQoutΔxΔy(8.10)

and

ΔθΔt=ΔQΔz.(8.11)

Equation (8.11) is the continuity equation for water, with the left-hand side the change in water storage and the right-hand side the flux divergence.

Figure 8.6 Water balance for a soil volume with the fluxes QinQin entering the volume and QoutQout exiting the volume.

In the notation of calculus, the continuity equation is

θt=Qz,(8.12)

and substituting Darcy’s law for QQ gives

θt=zKθψz+Kθ=zKθψz+Kz.(8.13)

This is the Richards equation and describes the movement of water in an unsaturated porous medium (Richards Reference Richards1931). Equation (8.13) is called the mixed-form equation because it includes the time rate of change in θθ on the left-hand side and the vertical gradient in ψψ on the right-hand side. Other forms of the Richards equation use the dependence between ψψ and θθ to express the equation in terms of only one unknown variable. The head-based, or ψψ-based, form transforms the storage term on the left-hand side of the equation from θθ to ψψ so that ψψ is the dependent variable. This uses the chain rule to expand ∂θ/∂tθ/t as

θt=ψt=Cψψt,(8.14)

in which C(ψ) = /Cψ=/ is known as the specific moisture capacity (m–1) and is the slope of the soil moisture retention curve. Then (8.13) is rewritten as

Cψψt=zKθψz+Kz.(8.15)

The ψψ-based form is applicable for unsaturated and saturated conditions and provides a continuous equation for water flow in the vadose zone and for groundwater. However, it is not mass conserving because the specific moisture capacity // that appears in the storage term itself depends on ψψ and so is not constant over a discrete time interval during which ψψ changes value (Milly Reference Milly1985; Celia et al. Reference Celia, Bouloutas and Zarba1990). Whereas (8.14) is mathematically correct, its temporal discretization over some time interval ΔtΔt (as required in numerical methods) is not equivalent. The moisture-based, or θθ-based, equation uses θθ as the dependent variable with

θt=zDθθz+Kz,(8.16)

in which D(θ) = K(θ)/C(ψ)Dθ=Kθ/Cψ is referred to as the hydraulic diffusivity (m2 s–1). In this equation, the specific moisture capacity appears within the spatial derivative. The θθ-based form is mass conserving but is restricted to the unsaturated zone because soil moisture does not vary within a saturated porous medium (soil moisture is bounded by 0 ≤ θ ≤ θsat0θθsat) whereas pressure head does vary. Furthermore, the equation is restricted to homogenous soils because θθ is not continuous across soil layers with different θ(ψ)θψ relationships. As a result, soils in which texture varies with depth have discontinuous vertical profiles of θθ, whereas ψψ is continuous even in inhomogeneous soils.

The Richards equation requires relationships for K(θ)Kθ and θ(ψ)θψ. Analytical solutions are difficult to obtain because these relationships are highly nonlinear. Instead, numerical methods are used, which requires first writing the finite difference approximation of the partial differential equation and then linearizing the nonlinear terms involving K(θ)Kθ and C(ψ)Cψ. The accuracy of the solution very much depends on the details of the numerical methods, the form of the Richards equation, and the size of the time step and grid spacing. The literature on numerical methods to solve the Richards equation is enormous. The next two sections provide an introduction to this literature with the caveat that many more numerical techniques are available and the merits of particular methods are still being debated.

Question 8.2 Soil at a depth of 5 cm has a matric potential of –478 mm, and the matric potential 50 cm deeper is –843 mm. Calculate the vertical water flux with a hydraulic conductivity of 2 mm h–1. What is the horizontal water flux if both locations are at the same depth in the soil but separated by 50 cm? Explain the difference between the two fluxes.

Question 8.3 In Darcy’s law given by (8.9), water flux has the units m H2O s–1, hydraulic conductivity is m s–1, and hydraulic head is m. Is Q=Kθzψρwatg+z an equivalent equation? What are the units for ψψ, K(θ)Kθ, and QQ in this equation? Derive the conversion factor for K(θ)Kθ from m s–1 to the same units as ψψ.

Question 8.4 A model calculates soil moisture in the unsaturated zone using the mixed-form Richards equation and solves the equation θt=zKθψzKz. Explain the difference between this equation and (8.13). Are the equations equivalent?

8.5 Finite Difference Approximation

The finite difference approximation for the Richards equation represents the soil as a network of discrete nodal points that vary in space and time similar to that for soil temperature. In doing so, it is necessary to remember that hydraulic conductivity is not constant but, rather, varies with depth depending on soil moisture so that the mixed-form equation is expanded as

θt=Kθ2ψz2+Kzψz+Kz.(8.17)

For reasons of numerical stability similar to soil temperature, (8.17) is solved using an implicit time discretization in which the spatial derivatives ∂K/∂zK/z, ∂ψ/∂zψ/z, and 2ψ/∂z22ψ/z2 are written numerically using a central difference approximation at time n + 1n+1, and the time derivative uses a backward difference approximation at n + 1n+1 (Appendix A4). For a vertical grid with discrete layers each equally spaced at a distance ΔzΔz and with zz positive in the upward direction so that layer ii is above layer i + 1i+1, the numerical form of (8.17) is

θin+1θinΔt=Kin+1Δz2ψi1n+12ψin+1+ψi+1n+1+Ki1n+1Ki+1n+12Δzψi1n+1ψi+1n+12Δz+Ki1n+1Ki+1n+12Δz.(8.18)

This is an implicit solution in which θθ and ψψ are expressed at n + 1n+1, and KK is similarly evaluated with θθ at n + 1n+1. Rearranging terms gives an equivalent form in which

θin+1θinΔt=Ki1/2n+1Δz2ψi1n+1ψin+1Ki+1/2n+1Δz2ψin+1ψi+1n+1+Ki1/2n+1Ki+1/2n+1Δz,(8.19)

with

Ki±1/2n+1=Kin+1±Ki+1n+1Ki1n+14.(8.20)

The ψψ-based form of the equation is obtained by replacing the left-hand side of (8.19) with Cin+1ψin+1ψin/Δt.

A more general derivation of (8.19) is obtained by considering the mass balance of a soil layer. Figure 8.7 depicts the soil profile in a cell-centered grid of NN discrete layers similar to that used for soil temperature. Soil layer ii has a thickness ΔziΔzi. Water content θiθi, matric potential ψiψi, and hydraulic conductivity KiKi are defined at the center of the layer at depth zizi and are uniform over the layer. Depth decreases in the downward direction from the surface so that z0 = 0z0=0 denotes the surface and zi + 1 < zizi+1<zi (i.e., depths are negative distances from the surface). The flux of water QiQi between adjacent layers ii and i + 1i+1 depends on the hydraulic conductivities of the two layers. The hydraulic conductivity Ki + 1/2Ki+1/2 replaces the vertically varying hydraulic conductivities in the two adjacent layers with an effective conductivity for an equivalent homogenous medium so that the flux QiQi is

Qi=Ki+1/2Δzi+1/2ψiψi+1Ki+1/2,(8.21)

with Δzi + 1/2 = zi − zi + 1Δzi+1/2=zizi+1 the distance between nodes ii and i + 1i+1. The flux Qi − 1Qi1 at the top of soil layer ii is similarly

Qi1=Ki1/2Δzi1/2ψi1ψiKi1/2,(8.22)

and Δzi − 1/2 = zi − 1 − ziΔzi1/2=zi1zi. With fluxes expressed for time n + 1n+1, the mass balance for soil layer ii is

θin+1θinΔt=Qi1n+1Qin+1Δzi(8.23)

so that

ΔziΔtθin+1θin=Ki1/2n+1Δzi1/2ψi1n+1ψin+1Ki+1/2n+1Δzi+1/2ψin+1ψi+1n+1+Ki1/2n+1Ki+1/2n+1.(8.24)

For constant soil layer thickness, this is equivalent to (8.19).

Figure 8.7 Multilayer soil water flow in a cell-centered grid oriented such that i = 1i=1 is the top soil layer at the surface, and i = Ni=N is the bottom soil layer. Each layer has a thickness ΔziΔzi. The depth zi − 1/2zi1/2 is the interface between adjacent layers i − 1i1 and ii, and zi + 1/2zi+1/2 is the interface between ii and i + 1i+1. Depths are negative distances from the surface so that Δzi = zi − 1/2 − zi + 1/2Δzi=zi1/2zi+1/2 (i.e., zi + 1/2 = zi − 1/2 − Δzizi+1/2=zi1/2Δzi). The depth zizi is defined at the center of layer ii so that zi = (zi − 1/2 + zi + 1/2)/2zi=zi1/2+zi+1/2/2. Δzi ± 1/2Δzi±1/2 is the grid spacing between ii and i ± 1i±1. Water content θiθi, matric potential ψiψi, and hydraulic conductivity KiKi are defined at the center of layer ii at depth zizi and are uniform over the layer. An effective hydraulic conductivity Ki ± 1/2Ki±1/2 is defined at the interface between soil layers at depth zi ± 1/2zi±1/2. Shown are (a) the first soil layer (i = 1i=1); (b) layers 1 < i < N1<i<N depicted generally as three soil layers denoted i − 1i1, ii, and i + 1i+1; and (c) the bottom soil layer (i = Ni=N). The surface matric potential ψ0ψ0 or the flux Q0Q0 provide the upper soil boundary condition, and the lower boundary condition at the bottom of the soil is ψN + 1ψN+1 or gravitational drainage QNQN.

Equation (8.24) describes the water balance of soil layers 1 < i < N1<i<N. Special equations are needed for the top (i = 1i=1) and bottom (i = Ni=N) layers to account for boundary conditions. Boundary conditions at the surface are specified in terms of ψ0ψ0 (Dirichlet boundary condition) or as a flux of water Q0Q0 into the soil (Neumann boundary condition). With the surface value ψ0ψ0 specified, (8.24) is still valid, in which case the corresponding flux of water into the soil is

Q0n+1=K1/2n+1Δz1/2ψ0n+1ψ1n+1K1/2n+1.(8.25)

Alternatively, Q0Q0 can be directly specified as the boundary condition (e.g., as an infiltration rate; negative into the soil), in which case the water balance for layer i = 1i=1 is

ΔziΔtθin+1θin=Q0n+1Ki+1/2n+1Δzi+1/2ψin+1ψi+1n+1Ki+1/2n+1.(8.26)

The lower boundary condition at the bottom of the soil column can similarly be specified in terms of ψψ or as a flux of water. When given as the matric potential ψN + 1ψN+1, the corresponding flux of water draining out of the soil column is

QNn+1=KN+1/2n+1ΔzN+1/2ψNn+1ψN+1n+1KN+1/2n+1.(8.27)

A common flux boundary condition specifies free drainage at the bottom of the soil column in which QN =  − KNQN=KN. This is referred to as a unit hydraulic gradient because ∂ψ/∂z = 0ψ/z=0 in the Darcian flux. The water balance of layer i = Ni=N is then

ΔziΔtθin+1θin=Ki1/2n+1Δzi1/2ψi1n+1ψin+1+Ki1/2n+1Kin+1.(8.28)

Alternatively, the soil column can be coupled with a groundwater model to allow for the influence of water table dynamics on soil moisture. The total change in water in the soil column equals the net flux into the soil so that conservation is given by

i=1Nθin+1θinΔzi=QNn+1Q0n+1Δt.(8.29)

Equation (8.20) defines the effective conductivity Ki ± 1/2Ki±1/2 from the central difference approximation for (∂K/∂z)(∂ψ/∂z)K/zψ/z. Hydraulic conductivity can differ substantially between layers because of vertical gradients in soil moisture (e.g., during infiltration into a dry soil). The equation used to represent the effective conductivity affects the accuracy of the numerical solution, and other expressions can be used for Ki ± 1/2Ki±1/2 (Haverkamp and Vauclin Reference Haverkamp and Vauclin1979; Warrick Reference Warrick1991, Reference Warrick2003). The effective conductivity can be defined as the arithmetic mean of the adjacent nodal conductivities in which

Ki ± 1/2 = 0.5(Ki + Ki ± 1)
Ki±1/2=0.5Ki+Ki±1.
(8.30)

Another expression is the geometric mean whereby

Ki ± 1/2 = (KiKi ± 1)1/2
Ki±1/2=KiKi±11/2.
(8.31)

The harmonic mean is the reciprocal of the arithmetic mean of the reciprocals with

Ki±1/2=10.5Ki1+Ki±11=2KiKi±1Ki+Ki±1.(8.32)

This latter equation is similar to (5.16) for the effective thermal conductivity and is obtained for constant ΔzΔz from continuity of flow across the interface so that the flux of water from depth zizi to zi + 1/2zi+1/2 equals the flux from zi + 1/2zi+1/2 to zi + 1zi+1. While the harmonic mean ensures continuity of fluxes at the interface, it is weighted towards the lower value of the two hydraulic conductivities and is the smallest of the three means. The arithmetic mean has the largest value, and the geometric mean has an intermediate value. The arithmetic mean is commonly used – e.g., as in the numerical solutions of Haverkamp et al. (Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977) and Celia et al. (Reference Celia, Bouloutas and Zarba1990).

The numerical form of the ψψ-based Richards equation is similar to (5.18) for soil temperature. For a soil with NN layers, this is a tridiagonal system of NN equations with NN unknown values of ψψ at time n + 1n+1. This is more obvious by rewriting the finite difference approximation of the mixed-form equation given by (8.24) in the ψψ-based form and rearranging terms to get

Ki1/2n+1Δzi1/2ψi1n+1+Cin+1ΔziΔt+Ki1/2n+1Δzi1/2+Ki+1/2n+1Δzi+1/2ψin+1Ki+1/2n+1Δzi+1/2ψi+1n+1=Cin+1ΔziΔtψin+Ki1/2n+1Ki+1/2n+1.(8.33)

A general form for this equation is

aiψi1n+1+biψin+1+ciψi+1n+1=di,(8.34)

or, in matrix notation (Appendix A6),

b1c10000a2b2c20000a3b3c300000000aN1bN1cN10000aNbN×ψ1n+1ψ2n+1ψ3n+1ψN1n+1ψNn+1=d1d2d3dN1dN,(8.35)

in which the matrix elements aa, bb, cc, and dd are evident from (8.33), with special values for i = 1i=1 and i = Ni=N to account for boundary conditions (Table 8.4).

Table 8.4 Tridiagonal terms for the ψψ-based Richards equation

Layeraiaibibicicididi
i = 1i=10Cin+1ΔziΔt+K1/2n+1Δz1/2ciKi+1/2n+1Δzi+1/2Cin+1ΔziΔtψin+K1/2n+1Δz1/2ψ0n+1+K1/2n+1Ki+1/2n+1
1 < i < N1<i<NKi1/2n+1Δzi1/2 Cin+1ΔziΔtaici Ki+1/2n+1Δzi+1/2 Cin+1ΔziΔtψin+Ki1/2n+1Ki+1/2n+1
i = Ni=NKi1/2n+1Δzi1/2Cin+1ΔziΔtai0Cin+1ΔziΔtψin+Ki1/2n+1KNn+1

Note: Boundary conditions are ψ0n+1 for the first layer (i = 1i=1) and free drainage for the bottom layer (i = Ni=N).

However, whereas soil temperature uses a linear equation, (8.33) is a nonlinear equation for ψψ because of the complex dependence of KK and CC on ψψ. This nonlinearity is evident when these terms are replaced with their various expressions given in Table 8.2. An additional complexity is that ψψ at n + 1n+1 also appears on the right-hand side of (8.33) through KK and CC. Solving a nonlinear equation is challenging, and solving the system of nonlinear equations required to represent NN soil layers is especially challenging. The solution requires linearizing (8.33) with respect to ψψ through various numerical methods. One simple linearization uses values of KK and CC obtained at the preceding time step (at time nn rather than n + 1n+1). This is an implicit solution for ψψ but with explicit linearization of KK and CC (Haverkamp et al. Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977).

A better numerical technique is the predictor–corrector method. This is a two-step solution that solves the Richards equation twice. The predictor step uses explicit linearization to solve for ψψ over one-half a full time step (Δt/2Δt/2) at time n + 1/2n+1/2 with KK and CC from time nn. The resulting values for ψψ are used to evaluate KK and CC at n + 1/2n+1/2, and the corrector step uses these to obtain ψψ over the full time step (ΔtΔt) at n + 1n+1. The method can be applied to the ψψ-based Richard equations (Haverkamp et al. Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977) or the θθ-based equation (Hornberger and Wiberg Reference Hornberger and Wiberg2005). These implementations use an implicit solution for the predictor step and the Crank–Nicolson method (Appendix A4) for the corrector step. The predictor equation for ψψ at n + 1/2n+1/2 is

Ki1/2nΔzi1/2ψi1n+1/2+CinΔziΔt/2+Ki1/2nΔzi1/2+Ki+1/2nΔzi+1/2ψin+1/2Ki+1/2nΔzi+1/2ψi+1n+1/2=CinΔziΔt/2ψin+Ki1/2nKi+1/2n.(8.36)

The corrector equation solves for ψψ over a full time step using the Crank–Nicolson method with fluxes evaluated at time nn and n + 1n+1 whereby

Ki1/2n+1/22Δzi1/2ψi1n+1+Cin+1/2ΔziΔt+Ki1/2n+1/22Δzi1/2+Ki+1/2n+1/22Δzi+1/2ψin+1Ki+1/2n+1/22Δzi+1/2ψi+1n+1=Cin+1/2ΔziΔtψin+Ki1/2n+1/22Δzi1/2ψi1nψinKi+1/2n+1/22Δzi+1/2ψinψi+1n+Ki1/2n+1/2Ki+1/2n+1/2.(8.37)

Equations (8.36) and (8.37) are both a tridiagonal system of linear equations and are easily solved for ψψ (Appendix A8).

Question 8.5 Soil temperature is commonly modeled with zero heat flux as the boundary condition at the bottom of the soil column. Explain how this is similar to the free drainage boundary condition for soil moisture.

Question 8.6 Compare the ψψ-based Richards equation given by (8.33) with that for soil temperature given by (5.19). What are the similarities? What is a key difference? Why are iterative methods required to solve the Richards equation but not soil temperature?

Question 8.7 Table 8.4 gives the tridiagonal coefficients aiai, bibi, cici, and didi for the ψψ-based Richards equation using the implicit method. Derive the same coefficients for the Crank–Nicolson method as used in the predictor–correct solution. What is the equation for the infiltration rate Q0Q0?

8.6 Iterative Numerical Solutions

Other numerical methods use iterative calculations. These methods approach the correct solution by using successive approximations in which values for KK and CC from one iteration are used at the next iteration. Picard iteration, which is an example of fixed-point iteration (Appendix A5), is one such numerical algorithm. As applied to the ψψ-based Richards equation, Picard iteration provides successive estimates for ψψ using values of KK and CC evaluated with the previous value of ψψ (Celia et al. Reference Celia, Bouloutas and Zarba1990). The iteration repeats until ψψ does not change value between iterations. With nn denoting time and mm denoting iteration, (8.33) is written as

Ki1/2n+1,mΔzi1/2ψi1n+1,m+1+Cin+1,mΔziΔt+Ki1/2n+1,mΔzi1/2+Ki+1/2n+1,mΔzi+1/2ψin+1,m+1Ki+1/2n+1,mΔzi+1/2ψi+1n+1,m+1=Cin+1,mΔziΔtψin+Ki1/2n+1,mKi+1/2n+1,m.(8.38)

The values of KK and CC are obtained from iteration mm so that (8.38) is a tridiagonal system of linear equations that is solved for ψψ at time n + 1n+1 and iteration m + 1m+1. It is more convenient to rewrite this equation to solve for the change in ψψ between iterations (δm + 1 = ψn + 1, m + 1 − ψn + 1, mδm+1=ψn+1,m+1ψn+1,m) rather than directly for ψψ itself. With this modification, (8.38) becomes

Ki1/2n+1,mΔzi1/2δi1m+1+Cin+1,mΔziΔt+Ki1/2n+1,mΔzi1/2+Ki+1/2n+1,mΔzi+1/2δim+1Ki+1/2n+1,mΔzi+1/2δi+1m+1=fin+1,m,(8.39)

with

fin+1,m=Ki1/2n+1,mΔzi1/2ψi1n+1,mψin+1,mKi+1/2n+1,mΔzi+1/2ψin+1,mψi+1n+1,m+Ki1/2n+1,mKi+1/2n+1,mCin+1,mΔziΔtψin+1,mψin.(8.40)

Terms are known at iteration mm, and (8.39) is solved for δδ at iteration m + 1m+1. The right-hand side of (8.39) is the ψψ-based Richards equation evaluated at the mmth iteration. As the solution converges, δδ becomes small so that the left-hand side approaches zero, and (8.39) is the standard finite difference approximation with KK and CC expressed for n + 1n+1. Picard iteration is a simple procedure that evaluates KK and CC for the current estimate of ψψ, uses these to solve for a new value of ψψ, and repeats this calculation until convergence is achieved. However, convergence can require many iterations and is not always guaranteed.

A more complex numerical method uses Newton–Raphson iteration to linearize the system of NN equations (Appendix A9). This method reformulates the solution in terms of finding the roots of the system of equations. Newton–Raphson iteration defines δm + 1δm+1 as given previously but uses a Taylor series approximation to linearize the Richards equation and solves for δm + 1δm+1 that satisfies the equation

fiψi1δi1m+1+fiψiδim+1+fiψi+1δi+1m+1=fin+1,m.(8.41)

The right-hand side is the Richards equation evaluated at iteration mm as in (8.40), and the left-hand side uses the partial derivatives ∂f/∂ψf/ψ evaluated at iteration mm. The iteration proceeds until δδ is less than some convergence criterion. Equation (8.41) is similar to Picard iteration; but whereas that method uses the standard terms in the Richards equation, Newton–Raphson iteration requires evaluating the partial derivatives with respect to ψψ. In linear algebra, the partial derivatives are referred to as the Jacobian matrix. The two methods differ in the computational efficiency and robustness of the numerical solution (Paniconi et al. Reference Paniconi, Aldama and Wood1991; Paniconi and Putti Reference Paniconi and Putti1994; Lehmann and Ackerer Reference Lehmann and Ackerer1998). Picard iteration may fail to converge or may need many iterations to converge. Newton–Raphson iteration requires evaluating a matrix of partial derivatives (the Jacobian) and is computationally more expensive per iteration but can converge in fewer iterations and provide a more robust solution (though it, too, can fail to converge).

The difficulty in using the ψψ-based Richards equation is that it does not conserve mass because the specific moisture capacity C(ψ)Cψ that appears in the water storage term is not constant over a time step (Milly Reference Milly1985; Celia et al. Reference Celia, Bouloutas and Zarba1990). Celia et al. (Reference Celia, Bouloutas and Zarba1990) devised a mass-conserving numerical solution for the mixed-form Richards equation that is a modified Picard iteration. The mixed-form equation is

ΔziΔtθin+1,m+1θin=Ki1/2n+1,mΔzi1/2ψi1n+1,m+1ψin+1,m+1Ki+1/2n+1,mΔzi+1/2ψin+1,m+1ψi+1n+1,m+1+Ki1/2n+1,mKi+1/2n+1,m,(8.42)

with nn referring to time and mm to iteration as before. Mass conservation is achieved by using a Taylor series approximation (Appendix A1) for θin+1,m+1 in which

θin+1,m+1=θin+1,m+Cin+1,mψin+1,m+1ψin+1,m,(8.43)

with

Cin+1,m=dθidψin+1,m.(8.44)

Substituting this expression into (8.42) and converting to residual form gives

Ki1/2n+1,mΔzi1/2δi1m+1+Cin+1,mΔziΔt+Ki1/2n+1,mΔzi1/2+Ki+1/2n+1,mΔzi+1/2δim+1Ki+1/2n+1,mΔzi+1/2δi+1m+1=fin+1,m,(8.45)

as with the ψψ-based equation, but now with

fin+1,m=Ki1/2n+1,mΔzi1/2ψi1n+1,mψin+1,mKi+1/2n+1,mΔzi+1/2ψin+1,mψi+1n+1,m+Ki1/2n+1,mKi+1/2n+1,mΔziΔtθin+1,mθin.(8.46)

Equation (8.45) is a tridiagonal system of linear equations that is solved for δδ. Table 8.5 gives the various terms in the tridiagonal equations. The iteration is repeated until some convergence criterion is satisfied. This can be an absolute threshold in which δi ∣  ≤ εaδiεa for each layer or can also include both an absolute and a relative error term in which case δiεa+εrψin+1,m. As the iteration converges, δiδi approaches zero and (8.45) reduces to the mixed-form Richards equation. The key difference compared with the ψψ-based Picard iteration is the use of a Taylor series approximation for θθ, and this ensures mass conservation. At convergence, ψin+1,m+1ψin+1,m approaches zero, thereby eliminating inaccuracy in evaluating CiCi.

Table 8.5 Tridiagonal terms for the modified Picard iteration of the mixed-form Richards equation

Layeraiaibibicicididi
i = 1i=10Cin+1,mΔziΔtciKi+1/2n+1,mΔzi+1/2K1/2n+1,mΔz1/2ψ0n+1ψin+1,mKi+1/2n+1,mΔzi+1/2ψin+1,mψi+1n+1,m+K1/2n+1,mKi+1/2n+1,mΔziΔtθin+1,mθin
1 < i < N1<i<NKi1/2n+1,mΔzi1/2 Cin+1,mΔziΔtaici Ki+1/2n+1,mΔzi+1/2 Ki1/2n+1,mΔzi1/2ψi1n+1,mψin+1,mKi+1/2n+1,mΔzi+1/2ψin+1,mψi+1n+1,m+Ki1/2n+1,mKi+1/2n+1,mΔziΔtθin+1,mθin
i = Ni=NKi1/2n+1,mΔzi1/2 Cin+1,mΔziΔtai 0Ki1/2n+1,mΔzi1/2ψi1n+1,mψin+1,m+Ki1/2n+1,mKNn+1,mΔziΔtθin+1,mθin

Note: Boundary conditions are ψ0n+1 for the first layer (i = 1i=1) and free drainage for the bottom layer (i = Ni=N).

Accurate solution of the Richards equation requires a small time step ΔtΔt and spatial increment ΔzΔz. This is particularly true during infiltration into dry soil, where there is a sharp wetting front. Some models utilize adaptive time stepping in which ΔtΔt is dynamically adjusted and varies between some minimum and maximum value based on specified criteria. One simple method is to adjust ΔtΔt at every time, based on the number of iterations required for convergence at the previous time (Paniconi et al. Reference Paniconi, Aldama and Wood1991; Paniconi and Putti Reference Paniconi and Putti1994). The time step is likely to be too short if few iterations are needed to achieve convergence, but is likely to be too long if convergence requires many iterations. The time step is increased by a specified factor if convergence is achieved in fewer than some number of iterations, is decreased if some number of iterations is exceeded, or is otherwise left unchanged. If the solution fails to converge after a maximum number of iterations, ΔtΔt is decreased by some fraction and the iteration is restarted.

Global models must simulate tens of thousands of soil columns over hundreds of years, and small vertical or temporal step sizes pose a large computational burden. In these models, a linear form of the θθ-based Richards equation is commonly used because it conserves mass for all step sizes. The linearization is attained using a Taylor series approximation and can be solved directly for θθ at n + 1n+1 without iteration or also with Newton–Raphson iteration. The cost with larger step sizes and fewer iterations is less accuracy in the solution. With reference to Figure 8.7b, the water balance for soil layer ii at time n + 1n+1 and iteration m + 1m+1 is

ΔziΔtθin+1,m+1θin=Qi1n+1,m+1+Qin+1,m+1.(8.47)

The flux Qi − 1Qi1 is linearized as

Qi1n+1,m+1=Qi1n+1,m+Qi1θi1δi1m+1+Qi1θiδim+1,(8.48)

and QiQi is

Qin+1,m+1=Qin+1,m+Qiθiδim+1+Qiθi+1δi+1m+1,(8.49)

with δm + 1 = θn + 1, m + 1 − θn + 1, mδm+1=θn+1,m+1θn+1,m. Substituting these expressions into (8.47), the water balance is

Qi1θi1δi1m+1ΔziΔt+Qi1θiQiθiδim+1+Qiθi+1δi+1m+1=Qi1n+1,mQin+1,m+ΔziΔtθin+1,mθin.(8.50)

The solution becomes more accurate with multiple iterations as δδ approaches zero. The complexity lies in the partial derivatives, which include expressions for ∂ψ/∂θψ/θ and ∂K/∂θK/θ.

Question 8.8 Contrast the predictor–corrector, modified Picard, and Newton–Raphson methods to solve the Richards equations. What are the main differences among these methods? What are the similarities?

Question 8.9 Use Newton–Raphson iteration to solve the system of equations: x12+x22=4 and x1x2 = 1x1x2=1.

8.7 Infiltration

The Richards equation can be used to model infiltration into soil. The boundary condition at the soil surface depends on the rate at which water is applied to the soil. When the supply rate is less than the saturated hydraulic conductivity, no water accumulates on the surface and a flux (Neumann) boundary condition is used. If sufficient water is provided so that the soil surface is saturated but water does not pond, a concentration (Dirichlet) boundary condition is used with θ0 = θsatθ0=θsat. If water ponds on the soil surface, the boundary condition is θ0 = θsatθ0=θsat and with a small positive depth of water on the surface. Figure 8.8 shows soil moisture profiles during infiltration into sand and Yolo light clay using the predictor–corrector method, as in Haverkamp et al. (Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977). Initial conditions for sand are θ = 0.1θ=0.1 at t = 0t=0, and the boundary condition is θ0 = 0.267θ0=0.267 for t>0t>0. The simulated soil moisture closely matches the analytical solution. For clay, the initial and boundary conditions are θ = 0.24θ=0.24 and θ0 = 0.495θ0=0.495. Infiltration into the clay proceeds much slower than the sand. The sand absorbs almost 12 cm of water in 48 minutes while 18 cm of water infiltrates into the clay over 11.6 days (Figure 8.9).

Figure 8.8 Soil moisture profiles for (a) sand and (b) Yolo light clay during infiltration with a specified θ0θ0 boundary condition. Simulations are as in Haverkamp et al. (Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977) and use the predictor–correction method. Results are shown at various times up to 0.8 h (48 minutes) for sand and 277.8 h (11.6 days) for clay. The open circles show the analytical solution from Haverkamp et al. (Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977).

Figure 8.9 As in Figure 8.8, but for cumulative infiltration.

8.8 Source and Sink Terms

The Richards equation, as discussed in this chapter, considers only the vertical Darcian fluxes of water. An additional source or sink term can be added (source) or subtracted (sink) to account for other water fluxes. Primary among these is evapotranspiration loss, which is accounted for by subtracting a root extraction, or sink, term in the Richards equation. Other plant-mediated water fluxes such as hydraulic redistribution can also be considered. In this case, the water balance is

ΔziΔtθin+1θin=Qi1n+1+Qin+1±Sw,i,(8.51)

where Sw, iSw,i (m s–1) is the flux of water added or subtracted in soil layer ii.

Evapotranspiration must be partitioned to root uptake in each soil layer. A common method uses the root profile weighted by a soil wetness factor βwβw. For soil layer ii, a simple wetness factor is

βw,i=ψiψdryψoptψdryψi>ψdry0ψiψdry,(8.52)

in which ψdryψdry is the matric potential at which transpiration ceases and ψoptψopt is the matric potential at which βw = 1βw=1. Some models use volumetric moisture rather than matric potential; the difference relates to the nonlinearity of the θ(ψ)θψ relationship. Total evapotranspiration EE is partitioned to an individual soil layer in relation to the relative root fraction ΔFiΔFi obtained from (2.23), and

Sw,i=EΔFiβw,i/i=1NΔFiβw,i.(8.53)

More complex models of plant hydraulics calculate root uptake from physiological principles (Chapter 13).

Hydraulic redistribution is a process by which roots move water upward and downward in the soil. At night, roots can transport water from wet, deep soil layers to dry, upper soil layers, thereby increasing the supply of water available to near-surface roots. Downward root-mediated transport can also occur from wet, upper layers to dry, lower layers after rainfall. Modeling studies show that hydraulic redistribution helps sustain photosynthesis and transpiration during the dry season (Lee et al. Reference Lee, Oliveira, Dawson and Fung2005; Zheng and Wang Reference Zheng and Wang2007; Baker et al. Reference Baker, Prihodko and Denning2008; Wang Reference Wang2011; Li et al. Reference Li, Wang and Yu2012b; Yan and Dickinson Reference Yan and Dickinson2014). Many models include the water flux from hydraulic redistribution as a source or sink term in the Richards equation based on Ryel et al. (Reference Ryel, Caldwell, Yoder, Or and Leffler2002), though Amenu and Kumar (Reference Amenu and Kumar2008) provided an alternative parameterization. The flux is calculated from the difference in matric potential between two layers using Darcy’s law. At night, the flux of water by hydraulic redistribution from layer jj to layer ii is

Sw,ji=CRTψjψimaxcicjΔFiΔFj1ΔFj,(8.54)

where CRTCRT is a maximum radial soil–root conductance (m MPa–1 s–1) and ΔψΔψ is the difference in water potential (MPa) between the uptake and release soil layers. The term cici or cjcj is a factor that reduces conductance for soil moisture in the source or sink layer and is given by

ci=11+ψi/ψ50b,(8.55)

in which ψ50ψ50 is the matric potential at which conductance is reduced by 50%. Representative parameters are CRT = 0.097CRT=0.097 cm MPa–1 h–1, ψ50 =  − 1ψ50=1 MPa, and b = 3.22b=3.22 (Ryel et al. Reference Ryel, Caldwell, Yoder, Or and Leffler2002), as used also in some land surface models (Zheng and Wang Reference Zheng and Wang2007; Wang Reference Wang2011; Li et al. Reference Li, Wang and Yu2012b; Yan and Dickinson Reference Yan and Dickinson2014). The rightmost term in (8.54) accounts for root abundance. The denominator is 1 − ΔFj1ΔFj when θj>θiθj>θi, but 1 − ΔFi1ΔFi when θi>θjθi>θj. Models commonly do not allow hydraulic redistribution to the top soil layer. Otherwise, the soil surface is continually wetted, and excessive soil evaporation can occur.

Question 8.10 Write the mixed-form Richards equation (8.13) with a sink term. What are the units of SS?

8.9 Soil Heterogeneity

The Richards equation is commonly used in land surface models. It applies to homogenous soil columns in which soil hydraulic properties are horizontally uniform but can vary in the vertical dimension. Spatially homogenous soil columns are applicable for laboratory studies or at small scales, but field soils are, in fact, quite heterogeneous. Several texture classes can co-occur within a small footprint, hydraulic conductivity and specific moisture capacity can differ within a texture class, and the presence of macropores can alter water movement in soils. Soil heterogeneity can be addressed through stochastic methods applied to the Richards equation, such as treating hydraulic conductivity as a random variable (Gelhar Reference Gelhar1986; Milly Reference Milly1988). These methods parameterize spatial heterogeneity statistically rather than explicitly representing the variability. A goal of such parameterizations is to obtain not only the mean water flow and moisture profile, but also the variance. A secondary goal is to obtain effective hydraulic parameters at large scales for which the Richards equation can be used. One approach is to represent soil as independent columns that vary in hydraulic parameters such as KsatKsat, for which the Richards equation is solved. This characterizes soil heterogeneity by a series of independent, one-dimensional flow problems. The mean soil moisture and its variance are obtained from an ensemble of Monte Carlo simulations in which, for example, KsatKsat is drawn from a specified probability density function or by numerically integrating the solution over the distribution (Bresler and Dagan Reference Bresler and Dagan1983; Clapp et al. Reference Clapp, Hornberger and Cosby1983; Dagan and Bresler Reference Dagan and Bresler1983). Another method is to solve the Richards equation in a probabilistic treatment to obtain a probability distribution for soil moisture with a mean and variance. This approach formulates the Richards equation as a stochastic partial differential equation with hydraulic conductivity taken as a random variable (Yeh et al. Reference Yeh, Gelhar and Gutjahr1985a,Reference Yeh, Gelhar and Gutjahrb,Reference Yeh, Gelhar and Gutjahrc; Mantoglou and Gelhar Reference Mantoglou and Gelhar1987a,Reference Mantoglou and Gelharb,Reference Mantoglou and Gelharc; Chen et al. Reference Chen, Govindaraju and Kavvas1994). A recent example is Vrettas and Fung (Reference Vrettas and Fung2015), who applied the concept to a local watershed but also suggested its applicability to large-scale land surface models.

8.10 Supplemental Programs

8.1 Predictor–Corrector Solution for the ψψ -Based Richards Equation: This program implements the predictor–correction method given by (8.36) and (8.37). Boundary conditions are θ0θ0 and free drainage. The code uses either the Campbell (Reference Campbell1974) or van Genuchten (Reference van Genuchten1980) relationships for θ(ψ)θψ and K(θ)Kθ. Specific configurations match Haverkamp et al. (Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977) for sand and Yolo light clay as in Figure 8.8 and Figure 8.9. θ(ψ)θψ uses the van Genuchten relationships with θres = 0.075θres=0.075, θsat = 0.287θsat=0.287, α = 0.027α=0.027 cm–1, n = 3.96n=3.96, and m = 1m=1 (sand); and θres = 0.124θres=0.124, θsat = 0.495θsat=0.495, α = 0.026α=0.026 cm–1, n = 1.43n=1.43, and m = 0.3m=0.3 (clay). In Haverkamp et al. (Reference Haverkamp, Vauclin, Touma, Wierenga and Vachaud1977), K = KsatA/(A + |ψ|B)K=KsatA/A+ψB with Ksat = 34Ksat=34 cm h–1, A = 1.175 × 106A=1.175×106, and B = 4.74B=4.74 (sand); and Ksat = 0.0443Ksat=0.0443 cm h–1, A = 124.6A=124.6, and B = 1.77B=1.77 (clay).

8.2 Modified Picard Iteration for the Mixed-Form Richards Equation: This program is similar to the previous program but implements the modified Picard iteration (8.45) with Table 8.5. Critical parameters for the solution are the tolerance εaεa, which is the maximum allowable change in ψψ between iterations for convergence. This parameter determines the accuracy of the water balance.

8.11 Modeling Projects

  1. 1. Use the ψψ-based predictor–corrector method (Supplemental Program 8.1) to calculate the amount of water that infiltrates into a sandy loam. Compare results using the van Genuchten (Reference van Genuchten1980) and Campbell (Reference Campbell1974) relationships for θ(ψ)θψ and K(θ)Kθ with parameters from Table 8.3.

  2. 2. Repeat the previous problem, but using the modified Picard iteration (Supplemental Program 8.2). How do the results compare with the predictor–corrector method? What can be said about parameter uncertainty versus numerical methods?

Figure 0

Figure 8.1 Water flows in a soil column extending from the ground surface to the water table.

Figure 1

Figure 8.2 Soil water movement during infiltration into sand. Shown are the initial moisture profile (θ = 0.1θ=0.1) and profiles in increments of 0.1 h.

Adapted from Haverkamp et al. (1977)
Figure 2

Figure 8.3 Depiction of soil with total volume VV comprising soil solids VsVs, water VwVw, and air VaVa. The total volume of pore space is Vp = Vw + VaVp=Vw+Va. The total soil mass consists of soil solids with mass msms and water with mass mwmw.

Figure 3

Table 8.1 Various measures of soil water for a volume of soil with dimensions 10 cm × 10 cm × 10 cm that has a mass of 1.7 kg when wet, 1.45 kg when dry, and particle density ρs = 2650ρs=2650 kg m–3

Figure 4

Figure 8.4 Water content in relation to matric potential given here as suction (ψ−ψ).Shown is the van Genuchten (1980) θ(ψ)θψrelationship for Berino loamy fine sand (θres = 0.0286θres=0.0286, θsat = 0.3658θsat=0.3658, α = 0.028α=0.028 cm–1, n = 2.239n=2.239) and Glendale silty clay loam (θres = 0.106θres=0.106, θsat = 0.4686θsat=0.4686, α = 0.0104α=0.0104 cm–1, n = 1.3954n=1.3954) from Hills et al. (1989).

Figure 5

Table 8.2 Soil moisture retention and hydraulic conductivity functions

Figure 6

Table 8.3 Parameter values for the Campbell (1974) and van Genuchten (1980) θ(ψ)θψ and K(θ)Kθ functions arranged by soil texture

Source: Campbell (1974) parameters from Clapp and Hornberger (1978); van Genuchten (1980) parameters from Carsel and Parrish (1988) and Leij et al. (1996).
Figure 7

Figure 8.5 Hydraulic conductivity in relation to water content. Shown is the van Genuchten (1980) K(θ)Kθrelationship for Berino loamy fine sand and Glendale silty clay loam. Parameter values are as in Figure 8.4 and KsatKsat= 22.54 and 0.55 cm h–1, respectively.

(Hills et al. 1989)
Figure 8

Figure 8.6 Water balance for a soil volume with the fluxes QinQin entering the volume and QoutQout exiting the volume.

Figure 9

Figure 8.7 Multilayer soil water flow in a cell-centered grid oriented such that i = 1i=1 is the top soil layer at the surface, and i = Ni=N is the bottom soil layer. Each layer has a thickness ΔziΔzi. The depth zi − 1/2zi−1/2 is the interface between adjacent layers i − 1i−1 and ii, and zi + 1/2zi+1/2 is the interface between ii and i + 1i+1. Depths are negative distances from the surface so that Δzi = zi − 1/2 − zi + 1/2Δzi=zi−1/2−zi+1/2 (i.e., zi + 1/2 = zi − 1/2 − Δzizi+1/2=zi−1/2−Δzi). The depth zizi is defined at the center of layer ii so that zi = (zi − 1/2 + zi + 1/2)/2zi=zi−1/2+zi+1/2/2. Δzi ± 1/2Δzi±1/2 is the grid spacing between ii and i ± 1i±1. Water content θiθi, matric potential ψiψi, and hydraulic conductivity KiKi are defined at the center of layer ii at depth zizi and are uniform over the layer. An effective hydraulic conductivity Ki ± 1/2Ki±1/2 is defined at the interface between soil layers at depth zi ± 1/2zi±1/2. Shown are (a) the first soil layer (i = 1i=1); (b) layers 1 < i < N1 depicted generally as three soil layers denoted i − 1i−1, ii, and i + 1i+1; and (c) the bottom soil layer (i = Ni=N). The surface matric potential ψ0ψ0 or the flux Q0Q0 provide the upper soil boundary condition, and the lower boundary condition at the bottom of the soil is ψN + 1ψN+1 or gravitational drainage QNQN.

Figure 10

Table 8.4 Tridiagonal terms for theψψ-based Richards equation

Figure 11

Table 8.5 Tridiagonal terms for the modified Picard iteration of the mixed-form Richards equation

Figure 12

Figure 8.8 Soil moisture profiles for (a) sand and (b) Yolo light clay during infiltration with a specified θ0θ0 boundary condition. Simulations are as in Haverkamp et al. (1977) and use the predictor–correction method. Results are shown at various times up to 0.8 h (48 minutes) for sand and 277.8 h (11.6 days) for clay. The open circles show the analytical solution from Haverkamp et al. (1977).

Figure 13

Figure 8.9 As in Figure 8.8, but for cumulative infiltration.

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Soil Moisture
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Climate Change and Terrestrial Ecosystem Modeling
  • Online publication: 08 February 2019
  • Chapter DOI: https://doi.org/10.1017/9781107339217.009
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Soil Moisture
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Climate Change and Terrestrial Ecosystem Modeling
  • Online publication: 08 February 2019
  • Chapter DOI: https://doi.org/10.1017/9781107339217.009
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Soil Moisture
  • Gordon Bonan, National Center for Atmospheric Research, Boulder, Colorado
  • Book: Climate Change and Terrestrial Ecosystem Modeling
  • Online publication: 08 February 2019
  • Chapter DOI: https://doi.org/10.1017/9781107339217.009
Available formats
×