You are on page 1of 75

Chapter 7

Groundwater Flow

Groundwater is water which is stored in the soil and rock beneath the surface of the
Earth. It forms a fundamental constituent reservoir of the hydrological system, and it
is important because of its massive and long lived storage capacity. It is the resource
which provides drinking and irrigation water for crops, and increasingly in recent
decades it has become an unwilling recipient of toxic industrial and agricultural
waste. For all these reasons, the movement of groundwater is an important subject
of study.
Soil consists of very small grains of organic and inorganic matter, ranging in size
from millimetres to microns. Differently sized particles have different names. Particularly, we distinguish clay particles (size <2 microns) from silt particles (260 microns) and sand (60 microns to 1 mm). Coarser particles still are termed gravel.
Viewed at the large scale, soil thus forms a continuum which is granular at the
small scale, and which contains a certain fraction of pore space, as shown in Fig. 7.1.
The volume fraction of the soil (or sediment, or rock) which is occupied by the pore
space (or void space, or voidage) is called the porosity, and is commonly denoted
by the symbol ; sometimes other symbols are used, for example n, as in Chap. 5.
As we described in Chap. 6, soils are formed by the weathering of rocks, and
are specifically referred to as soils when they contain organic matter formed by the
rotting of plants and animals. There are two main types of rock: igneous, formed by
the crystallisation of molten lava, and sedimentary, formed by the cementation of
sediments under conditions of great temperature and pressure as they are buried at
depth.1 Sedimentary rocks, such as sandstone, chalk, shale, thus have their porosity
built in, because of the pre-existing granular structure. With increasing pressure,
the grains are compacted, thus reducing their porosity, and eventually intergranular
cements bond the grains into a rock. Sediment compaction is described in Sect. 7.11.
Igneous rock tends to be porous also, for a different reason. It is typically the
case for any rock that it is fractured. Most simply, rock at the surface of the Earth
1 There are also

metamorphic rocks, which form from pre-existing rocks through chemical changes
induced by burial at high temperatures and pressures; for example, marble is a metamorphic form
of limestone.

A. Fowler, Mathematical Geoscience, Interdisciplinary Applied Mathematics 36,


DOI 10.1007/978-0-85729-721-1_7, Springer-Verlag London Limited 2011

387

388

7 Groundwater Flow

Fig. 7.1 A granular porous


medium

is subjected to enormous tectonic stresses, which cause folding and fracturing of


rock. Thus, even if the rock matrix itself is not porous, there are commonly faults
and fractures within the rock which act as channels through which fluids may flow,
and which act on the large scale as an effective porosity. If the matrix is porous at
the grain scale also, then one refers to the rock as having a dual porosity, and the
corresponding flow models are called double porosity models.
In the subsurface, whether it be soil, underlying regolith, a sedimentary basin,
or oceanic lithosphere, the pore space contains liquid. At sufficient depth, the pore
space will be saturated with fluid, normally water. At greater depths, other fluids
may be present. For example, oil may be found in the pore space of the rocks of
sedimentary basins. In the near surface, both air and water will be present in the
pore space, and this (unsaturated) region is called the unsaturated zone, or the vadose
zone. The surface separating the two is called the piezometric surface, the phreatic
surface, or more simply the water table. Commonly it lies tens of metres below the
ground surface.

7.1 Darcys Law


Groundwater is fed by surface rainfall, and as with surface water it moves under a
pressure gradient driven by the slope of the piezometric surface. In order to characterise the flow of a liquid in a porous medium, we must therefore relate the flow
rate to the pressure gradient. An idealised case is to consider that the pores consist
of uniform cylindrical tubes of radius a; initially we will suppose that these are all
aligned in one direction. If a is small enough that the flow in the tubes is laminar
(this will be the case if the associated Reynolds number is 1000), then Poiseuille
4
flow in each tube leads to a volume flux in each tube of q = a
8 |p|, where is

7.1 Darcys Law

389

the liquid viscosity, and p is the pressure gradient along the tube. A more realistic
porous medium is isotropic, which is to say that if the pores have this tubular shape,
the tubules will be arranged randomly, and form an interconnected network. However, between nodes of this network, Poiseuille flow will still be appropriate, and an
appropriate generalisation is to suppose that the volume flux vector is given by
q

a4
p,
X

(7.1)

where the approximation takes account of small interactions at the nodes; the numerical tortuosity factor X  1 takes some account of the arrangement of the pipes.
To relate this to macroscopic variables, and in particular the porosity , we observe that a 2 /dp2 , where dp is a representative particle or grain size so that
2 d 2

q/dp2 ( Xp )p. We define the volume flux per unit area (having units of velocity) as the discharge u. Darcys law then relates this to an applied pressure gradient
by the relation
k
u = p,

(7.2)

where k is an empirically determined parameter called the permeability, having units


of length squared. The discussion above suggests that we can write
k=

dp2 2
X

(7.3)

the numerical factor X may typically be of the order of 103 .


To check whether the pore flow is indeed laminar, we calculate the (particle)
Reynolds number for the porous flow. If v is the (average) fluid velocity in the pore
space, then
v=

u
.

(7.4)

If a is the pore radius, then we define a particle Reynolds number based on grain
size as
Rep =

2va |u|dp
,

(7.5)

since a/dp . Suppose (7.3) gives the permeability, and we use the gravitational
pressure gradient g to define (via Darcys law) a velocity scale2 ; then
 

3/2 gdp dp 2
10[dp ]3 ,
Rep
X

2 This

scale is thus the hydraulic conductivity, defined below in (7.9).

(7.6)

390

7 Groundwater Flow

where dp = [dp ] mm, and using 3/2 /X = 103 , g = 10 m s2 , / = 106 m2 s2 .


Thus the flow is laminar for d < 5 mm, corresponding to a gravel. Only for free flow
through very coarse gravel could the flow become turbulent, but for water percolation in rocks and soils, we invariably have slow, laminar flow.
In other situations, and notably for forced gas stream flow in fluidised beds or in
packed catalyst reactor beds, the flow can be rapid and turbulent. In this case, the
Poiseuille flow balance p = u/k can be replaced by the Ergun equation
p =

|u|u
;
k

(7.7)

more generally, the right hand side will be a sum of the two (laminar and turbulent)
interfacial resistances. The Ergun equation reflects the fact that turbulent flow in a
pipe is resisted by Reynolds stresses, which are generated by the fluctuation of the
inertial terms in the momentum equation. Just as for the laminar case, the parameter k  , having units of length, depends both on the grain size dp and on . Evidently,
we will have
k  = dp E(),

(7.8)

with the numerical factor E 0 as 0.

7.1.1 Hydraulic Conductivity


Another measure of flow rate in porous soil or rock relates specifically to the passage
of water through a porous medium under gravity. For free flow, the pressure gradient
downwards due to gravity is just g, where is the density of water and g is the
gravitational acceleration; thus the water flux per unit area in this case is just
K=

kg
,

(7.9)

and this quantity is called the hydraulic conductivity. It has units of velocity. A hydraulic conductivity of K = 105 m s1 (about 300 m y1 ) corresponds to a permeability of k = 1012 m2 , this latter unit also being called the darcy.

7.1.2 Homogenisation
The derivation of Darcys law can be carried out in a more formal way using the
method of homogenisation. This is essentially an application of the method of multiple (space) scales to problems with microstructure. Usually (for analytic reasons)
one assumes that the microstructure is periodic, although this is probably not strictly
necessary (so long as local averages can be defined).

7.1 Darcys Law

391

Consider the Stokes flow equations for a viscous fluid in a medium of macroscopic length l, subject to a pressure gradient of order p/ l. If the microscopic
(e.g., grain size) length scale is dp , and = dp / l, then if we scale velocity with
dp2 p/ l (appropriate for local Poiseuille-type flow), length with l, and pressure
with p, the NavierStokes equations can be written in the dimensionless form
.u = 0,
0 = p + 2 2 u,

(7.10)

together with the no-slip boundary condition,


u = 0 on S : f (x/) = 0,

(7.11)

where S is the interfacial surface. We put x = and seek solutions in the form
u = u(0) (x, ) + u(1) (x, ) . . . ,
p = p(0) (x, ) + p (1) (x, ) . . . .

(7.12)

Expanding the equations in powers of and equating terms leads to p(0) = p (0) (x),
and u(0) satisfies
.u(0) = 0,
0 = p(1) + 2 u(0) x p (0) ,

(7.13)

equivalent to Stokes equations for u(0) with a forcing term x p (0) . If wj is the
velocity field which (uniquely) solves
.wj = 0,
0 = P + 2 wj + ej ,

(7.14)

with periodic (in ) boundary conditions and u = 0 on f ( ) = 0, where ej is the


unit vector in the j direction, then (since the equation is linear) we have (summing
over j )3
u(0) =

p (0) j
w .
xj

(7.15)

We define the average flux


u =

3 In

1
V


u(0) dV ,

(7.16)

other words, we employ the summation convention which states that summation is implied
over repeated suffixes, see for example Jeffreys and Jeffreys (1953).

392

7 Groundwater Flow

where V is the volume over which S is periodic.4 Averaging (7.15) then gives
u = k .p,
where the (dimensionless) permeability tensor is defined by
 j
kij = wi .

(7.17)

(7.18)

Recollecting the scales for velocity, length and pressure, we find that the dimensional version of (7.17) is
k
u = .p,

(7.19)

k = k dp2 ,

(7.20)

where
so that k is the equivalent in homogenisation theory of the quantity 2 /X in (7.3).

7.1.3 Empirical Measures


While the validity of Darcys law can be motivated theoretically, it ultimately relies
on experimental measurements for its accuracy. The permeability k has dimensions
of (length)2 , which as we have seen is related to the mean grain size. If we write
k = dp2 C, then the number C depends on the pore configuration. For a tubular network (in three dimensions), one finds C 2 /72 (as long as is relatively small).
A different and often used relation is that of Carman and Kozeny, which applies to
pseudo-spherical grains (for example sand grains); this is
C

3
.
180(1 )2

(7.21)

The factor (1 )2 takes some account of the fact that as increases towards one,
the resistance to motion becomes negligible. In fact, for media consisting of uncemented (i.e., separate) grains, there is a critical value of beyond which the medium
as a whole will deform like a fluid. Depending on the grain size distribution, this
value is about 0.5 to 0.6. When the medium deforms in this way, the description
of the intergranular fluid flow can still be taken to be given by Darcys law, but
this now constitutes a particular choice of the interactive drag term in a two-phase
flow model. At lower porosities, deformation can still occur, but it is elastic not viscous (on short time scales), and given by the theory of consolidation or compaction,
which we discuss later.
4 Specifically,

we take V to be the soil volume, but the integral is only over the pore space volume,
where u is defined. In that case, the average u is in fact the Darcy flux (i.e., volume fluid flux per
unit area).

7.2 Basic Groundwater Flow


Table 7.1 Different grain
size materials and their
typical permeabilities

393
k (m2 )

Material

108

gravel

1010

sand

1012

fractured igneous rock

1013

sandstone

1014

silt

1018

clay

1020

granite

In the case of soils or sediments, empirical power laws of the form


C m

(7.22)

are often used, with much higher values of the exponent (e.g. m = 8). Such behaviour reflects the (chemically derived) ability of clay-rich soils to retain a high
fraction of water, thus making flow difficult. Table 7.1 gives typical values of the
permeability of several common rock and soil types, ranging from coarse gravel
and sand to finer silt and clay.
An explicit formula of CarmanKozeny type for the turbulent Ergun equation
expresses the turbulent permeability k  , defined in (7.7), as
k =

3 dp
.
175(1 )

(7.23)

7.2 Basic Groundwater Flow


Darcys equation is supplemented by an equation for the conservation of the fluid
phase (or phases, for example in oil recovery, where these may be oil and water).
For a single phase, this equation is of the simple conservation form

() + .(u) = 0,
t

(7.24)

supposing there are no sources or sinks within the medium. In this equation, is the
material density, that is, mass per unit volume of the fluid. A term is not present
in the divergence term, since u has already been written as a volume flux (i.e., the
has already been included in it: cf. (7.4)).
Eliminating u, we have the parabolic equation


k
() = . p ,
(7.25)
t

and we need a further equation of state (or two) to complete the model. The simplest
assumption corresponds to incompressible groundwater flowing through a rigid

394

7 Groundwater Flow

porous medium. In this case, and are constant, and the governing equation
reduces (if also k is constant) to Laplaces equation
2 p = 0.

(7.26)

This simple equation forms the basis for the following development. Before pursuing this, we briefly mention one variant, and that is when there is a compressible
pore fluid (e.g., a gas) in a non-deformable medium. Then is constant (so k is constant), but is determined by pressure and temperature. If we can ignore the effects
of temperature, then we can assume p = p() with p () > 0, and
t =

k
. p  () ,

(7.27)

which is a nonlinear diffusion equation for , sometimes called the porous medium
equation. If p , > 0, this is degenerate when = 0, and the solutions display
the typical feature of finite spreading rate of compactly supported initial data.

7.2.1 Boundary Conditions


The Laplace equation (7.26) in a domain D requires boundary data to be prescribed
on the boundary D of the spatial domain. Typical conditions which apply are a no
flow through condition at an impermeable boundary, u.n = 0, whence
p
= 0 on D,
n

(7.28)

p = pa

(7.29)

or a permeable surface condition


on D,

where for example pa would be atmospheric pressure at the ground surface. Another
example of such a condition would be the prescription of oceanic pressure at the
interface with the oceanic crust.
A more common application of the condition (7.29) is in the consideration of
flow in the saturated zone below the water table (which demarcates the upper limit
of the saturated zone). At the water table, the pressure is in equilibrium with the air
in the unsaturated zone, and (7.29) applies. The water table is a free surface, and
an extra kinematic condition is prescribed to locate it. This condition says that the
phreatic surface is also a material surface for the underlying groundwater flow, so
that its velocity is equal to the average fluid velocity (not the flux): bearing in mind
(7.4), we have
F
u
+ .F = 0 on D,
t

if the free surface D is defined by F (x, t) = 0.

(7.30)

7.2 Basic Groundwater Flow

395

7.2.2 Dupuit Approximation


One of the principally obvious features of mature topography is that it is relatively
flat. A slope of 0.1 is very steep, for example. As a consequence of this, it is typically also the case that gradients of the free groundwater (phreatic) surface are also
small, and a consequence of this is that we can make an approximation to the equations of groundwater flow which is analogous to that used in shallow water theory
or the lubrication approximation, i.e., we can take advantage of the large aspect ratio of the flow. This approximation is called the Dupuit, or DupuitForchheimer,
approximation.
To be specific, suppose that we have to solve
2p = 0

in 0 < z < h(x, y, t),

(7.31)

where z is the vertical coordinate, z = h is the phreatic surface, and z = 0 is


an impermeable basement. We let u denote the horizontal (vector) component of
the Darcy flux, and w the vertical component. In addition, we now denote by

= ( x
, y
) the horizontal component of the gradient vector. The boundary conditions are then
p = 0,

ht + u.h = w

p
+ g = 0
z

on z = h,
(7.32)

on z = 0;

here we take (gauge) pressure measured relative to atmospheric pressure. The condition at z = 0 is that of no normal flux, allowing for gravity.
Let us suppose that a horizontal length scale of relevance is l, and that the corresponding variation in h is of order d, thus
=

d
l

(7.33)

is the size of the phreatic gradient, and is small. We non-dimensionalise the variables
by scaling as follows:
x, y l,
u

z d,

kgd
,
l

p gd,
kgd 2
,
l 2

l 2
.
kgd

(7.34)

The choice of scales is motivated by the same ideas as lubrication theory. The pressure is nearly hydrostatic, and the flow is nearly horizontal.
The dimensionless equations are
u = p,
.u + wz = 0,

2 w = (pz + 1),

(7.35)

396

7 Groundwater Flow

with
pz = 1 on z = 0,
p = 0,

ht = w + p.h

on z = h.

At leading order as 0, the pressure is hydrostatic:



p = h z + O 2 .

(7.36)

(7.37)

More precisely, if we put


p = h z + 2 p1 + ,

(7.38)

p1zz = 2 h,

(7.39)

then (7.35) implies

with boundary conditions, from (7.36),


p1z = 0 on z = 0,
p1z = ht + |h|2

on z = h.

(7.40)

Integrating (7.39) from z = 0 to z = h thus yields the evolution equation for h in the
form
ht = .[hh],

(7.41)

which is a nonlinear diffusion equation of degenerate type when h = 0.


This is easily solved numerically, and there are various exact solutions which
are indicated in the exercises. In particular, steady solutions are found by solving
Laplaces equation for 12 h2 , and there are various kinds of similarity solution. (7.41)
is a second order equation requiring two boundary conditions. A typical situation in
a river catchment is where there is drainage from a watershed to a river. A suitable
problem in two dimensions is
ht = (hhx )x + r,

(7.42)

where the source term r represents recharge due to rainfall. It is given by


r=

rD
,
2 K

(7.43)

where rD is the rainfall rate and K = kg/ is the hydraulic conductivity. At the
divide (say, x = 0), we have hx = 0, whereas at the river (say, x = 1), the elevation
is prescribed, h = 1 for example. The steady solution is

1/2
h = 1 + r rx 2
,

(7.44)

and perturbations to this decay exponentially. If this value of the elevation of the
water table exceeds that of the land surface, then a seepage face occurs, where water

7.2 Basic Groundwater Flow

397

seeps from below and flows over the surface. This can sometimes be seen in steep
mountainous terrain, or on beaches, when the tide is going out.
The Dupuit approximation is not uniformly valid at x = 1, where conditions of
symmetry at the base of a valley would imply that u = 0, and thus px = 0. There is
therefore a boundary layer near x = 1, where we rescale the variables by writing
x = 1 X,

w=

W
,

h = 1 + H,

p = 1 z + P .

(7.45)

Substituting these into the two-dimensional version of (7.35) and (7.36), we find
u = PX ,

W = Pz ,

2P = 0

in 0 < z < 1 + H, 0 < X < , (7.46)

with boundary conditions


P = H,

Ht + PX HX =

W
+r

on z = 1 + H,

PX = 0 on X = 0,

(7.47)

Pz = 0 on z = 0,
P H rX

as X .

At leading order in , this is simply


2P = 0

in 0 < z < 1, 0 < X < ,

Pz = 0 on z = 0, 1,

(7.48)

PX = 0 on X = 0,
P rX

as X .

Evidently, this has no solution unless we allow the incoming groundwater flux r
from infinity to drain to the river at X = 0, z = 1. We do this by having a singularity
in the form of a sink at the river,
P


r  2
ln X + (1 z)2

near X = 0, z = 1.

(7.49)

The solution to (7.48) can be obtained by using complex variables and the method
of images, by placing sinks at z = (2n + 1), for integral values of n. Making use
of the infinite product formula (Jeffrey 2004, p. 72)


1+
1

2
(2n + 1)2


= cosh

,
2

where = X + iz, we find the solution to be




X
z
X 2 z
r
cos2
+ sinh2
sin
.
P = ln cosh2

2
2
2
2

(7.50)

(7.51)

398

7 Groundwater Flow

Fig. 7.2 Groundwater flow


lines towards a river at X = 0,
z=1

The complex variable form of the solution is


= P + i =

2r

ln cosh
,

(7.52)

which is convenient for plotting. The streamlines of the flow are the lines =
constant, and these are shown in Fig. 7.2.
This figure illustrates an important point, which is that although the flow towards
a drainage point may be more or less horizontal, near the river the groundwater
seeps upwards from depth. Drainage is not simply a matter of near surface recharge
and drainage. This means that contaminants which enter the deep groundwater may
reside there for a very long time.
A related point concerns the recharge parameter r defined in (7.43). According to
Table 7.1, a typical permeability for sand is 1010 m2 , corresponding to a hydraulic
conductivity of K = 103 m s1 , or 3 104 m y1 . Even for phreatic slopes as
low as = 102 , the recharge parameter r  O(1), and shallow aquifer drainage is
feasible.
However, finer-grained sediments are less permeable, and the calculation of r
for a silt with permeability of 1014 m2 (K = 107 m s1 = 3 m y1 suggests that
r 1/2 1, so that if the Dupuit approximation applied, the groundwater surface
would lie above the Earths surface everywhere. This simply points out the obvious fact that if the groundmass is insufficiently permeable, drainage cannot occur
through it but water will accumulate at the surface and drain by overland flow. The
fact that usually the water table is below but quite near the surface suggests that the
long term response of landscape to recharge is to form topographic gradients and
sufficiently deep sedimentary basins so that this status quo can be maintained.

7.3 Unsaturated Soils


Let us now consider flow in the unsaturated zone. Above the water table, water and
air occupy the pore space. If the porosity is and the water volume fraction per
unit volume of soil is W , then the ratio S = W/ is called the relative saturation. If

7.3 Unsaturated Soils

399

Fig. 7.3 Configuration of air and water in pore space. The contact angle measured through the
water is acute, so that water is the wetting phase. ws , as and aw are the surface energies of the
three interfaces

S = 1, the soil is saturated, and if S < 1 it is unsaturated. The pore space of an unsaturated soil is configured as shown in Fig. 7.3. In particular, the air/water interface
is curved, and in an equilibrium configuration the curvature of this interface will
be constant throughout the pore space. The value of the curvature depends on the
amount of liquid present. The less liquid there is (i.e., the smaller the value of S),
then the smaller the pores where the liquid is found, and thus the higher the curvature. Associated with the curvature is a suction effect due to surface tension across
the air/water interface. The upshot of all this is that the air and water pressures are
related by a capillary suction characteristic function which expresses the difference
between the pressures as a function of mean curvature, and hence, directly, S:
pa pw = f (S).

(7.53)

The suction characteristic f (S) is equal to 2 , where is the mean interfacial


curvature: is the surface tension. For air and water in soil, f is positive as water
is the wetting phase, that is, the contact angle at the contact line between air, water
and soil grain is acute, measured through the water (see Fig. 7.3). The resulting form
of f (S) displays hysteresis as indicated in Fig. 7.4, with different curves depending
on whether drying or wetting is taking place.

7.3.1 The Richards Equation


To model unsaturated flow, we have the conservation of mass equation in the form
(S)
+ .u = 0,
t

(7.54)

where we take as constant. Darcys law for an unsaturated flow has the form, now
with gravitational acceleration included,
u=

k(S)

[p + g k],

(7.55)

400

7 Groundwater Flow

Fig. 7.4 Capillary suction


characteristic. It displays
hysteresis in wetting and
drying

where k is a unit vector upwards, and the permeability k depends on S. If k(1) = k0


(the saturated permeability), then one commonly writes k = k0 krw (S), where krw
is the relative permeability. The most obvious assumption would be krw = S, but
this is rarely appropriate, and a better representation is a convex function, such as
0 3
krw = S 3 . An even better representation is a function such as krw = ( SS
1S0 )+ , where
S0 is known as the residual saturation. It represents the fact that in fine-grained
soils, there is usually some minimal water fraction which cannot be removed. It is
naturally associated with a capillary suction characteristic function pa p = f (S)
which tends to infinity as S S0 +, also appropriate for fine-grained soils.
In one dimension, and if we take the vertical coordinate z to point downwards,
we obtain the Richards equation



S
f
k0

=
krw (S)
+ g .
(7.56)
t
z
z
We are assuming pa = constant (and also that the soil matrix is incompressible).

7.3.2 Non-dimensionalisation
We choose scales for the variables as follows:
f=

,
dp

,
gdp

z
,
gk0

(7.57)

where dp is grain size and is the surface tension, assumed constant. The Richards
equation then becomes, in dimensionless variables,

 
S

=
+1 .
(7.58)
krw
t
z
z
To be specific, we consider the case of soil wetting due to surface infiltration: of
rainfall, for example. Suitable boundary conditions for infiltration are
S = 1 at z = 0

(7.59)

7.3 Unsaturated Soils

401

if surface water is ponded, or





u0
u0
,
+ 1 = u =
=
krw
z
k0 w g K0

(7.60)

if there is a prescribed downward flux u0 ; K0 is the saturated hydraulic conductivity. In a dry soil we would have S 0 as z , or if there is a water table
at z = zp , S = 1 there.5 For silt with k0 = 1014 m2 , the hydraulic conductivity
K0 107 m s1 or 3 m y1 , while average rainfall in England, for example, is
1 m y1 . Thus on average u 1, but during storms we can expect u 1. For
large values of u , the desired solution may have S > 1 at z = 0; in this case ponding
occurs (as one observes), and (7.60) is replaced by (7.59), with the pond depth being
determined by the balance between accumulation, infiltration, and surface run-off.

7.3.3 Snow Melting


An application of the unsaturated flow model occurs in the study of melting snow.
In particular, it is found that pollutants which may be uniformly distributed in snow
(e.g. SO2 from sulphur emissions via acid rain) can be concentrated in melt water run-off, with a consequent enhanced detrimental effect on stream pollution. The
question then arises, why this should be so. We shall find that uniform surface melting of a dry snowpack can lead to a meltwater spike at depth.
Suppose we have a snow pack of depth d. Snow is a porous aggregate of ice
crystals, and meltwater formed at the surface can percolate through the snow pack
to the base, where run-off occurs. (We ignore effects of re-freezing of meltwater.)
The model (7.58) is appropriate, but the relevant length scale is d. Therefore we
define a parameter

,
(7.61)
=
gddp
and we rescale the variables as z 1/, t 1/. To be specific, we will also take
krw = S 3 ,

(7.62)

and
(S) =

1
S,
S

(7.63)

based on typical experimental results.


Suitable boundary conditions in a melting event might be to prescribe the melt
flux u0 at the surface, thus



u0
krw
at z = 0.
(7.64)
+ 1 = u =
z
K0
5 With

constant air pressure, continuity of S follows from continuity of pore water pressure.

402

7 Groundwater Flow

If the base is impermeable, then





krw
+ 1 = 0 at z = h.
z

(7.65)

This is certainly not realistic if S reaches 1 at the base, since then ponding must
occur and presumably melt drainage will occur via a channelised flow, but we examine the initial stages of the flow using (7.65). Finally, we suppose S = 0 at t = 0.
Again, this is not realistic in the model (it implies infinite capillary suction) but it is
a feasible approximation to make.
Simplification of this model now leads to the dimensionless DarcyRichards
equation in the form



S
2 S
2 S
+ 3S
=
S 1+S
.
(7.66)
t
z
z
z
If we choose = 70 mN m1 , dp = 0.1 mm, = 103 kg m3 , g = 10 m s2 ,
d = 1 m, then = 0.07. It follows that (7.66) has a propensity to form shocks,
these being diffused by the term in over a distance O() (by analogy with the
shock structure for the Burgers equation, see Chap. 1).
We want to solve (7.66) with the initial condition
S=0

at t = 0,

(7.67)

and the boundary conditions



S
= u
S 3 S 1 + S 2
z

on z = 0,

(7.68)

and

S
= 0 at z = 1.
S 3 S 1 + S 2
z
Roughly, for  1, these are
S = S0

at z = 0,

S = 0 at z = 1,

(7.69)

(7.70)

where S0 = u1/3 , which we initially take to be O(1) (and <1, so that surface ponding does not occur).
Neglecting , the solution is the step function
S = S0 ,
S = 0,

z < zf ,
z > zf ,

(7.71)

and the shock front at zf advances at a rate z f given by the jump condition
z f =

[S 3 ]+

[S]+

= S02 .

(7.72)

7.3 Unsaturated Soils

403

Fig. 7.5 S(Z) given by


(7.78); the shock front
terminates at the origin

In dimensional terms, the shock front moves at speed u0 /S0 , which is in fact obvious (given that it has constant S behind it).
The shock structure is similar to that of Burgers equation. We put
z = zf + Z,
and S rapidly approaches the quasi-steady solution S(Z) of



V S  + 3S 2 S  = S 1 + S 2 S  ,
where V = z f ; hence





S 1 + S 2 S  = S S02 S 2 ,

(7.73)

(7.74)

(7.75)

in order that S S0 as Z , and where we have chosen


V = S02 ,

(7.76)

(as S+ = 0), thus reproducing (7.72). The solution is a quadrature,




(1 + S 2 ) dS
= Z,
(S02 S 2 )

(7.77)

with an arbitrary added constant (amounting to an origin shift for Z). Hence


(1 + S02 )
S0 + S
ln
= Z.
S
2S0
S0 S

(7.78)

The shock structure is shown in Fig. 7.5; the profile terminates where S = 0 at
Z = 0. In fact, (7.75) implies that S = 0 or (7.78) applies. Thus when S given by
(7.78) reaches zero, the solution switches to S = 0. The fact that S/Z is discontinuous is not a problem because the diffusivity S(1 + S 2 ) goes to zero when S = 0.
This degeneracy of the equation is a signpost for fronts with discontinuous derivatives: essentially, the profile can maintain discontinuous gradients at S = 0 because
the diffusivity is zero there, and there is no mechanism to smooth the jump away.
Suppose now that k0 = 1010 m2 and / = 106 m2 s1 ; then the saturated hydraulic conductivity K0 = k0 g/ = 103 m s1 . On the other hand, if

404

7 Groundwater Flow

a metre thick snow pack melts in ten days, this implies u0 106 m s1 . Thus
S03 = u0 /K0 103 , and the approximation S S0 looks less realistic. With

S
= S03 ,
S 3 S 1 + S 2
z

(7.79)

and S0 101 and 101 , it seems that one should assume S  1. We define

S=

S03

1/2
s;

(7.80)

(7.79) becomes


S3
s
s 3 s 1 + 0 s 2
= 1 on z = 0,

(7.81)

and we have S03 / 102 , = (S0 /)3/2 1.


We neglect the term in S03 /, so that
s 3 s

s
1 on z = 0,
z

(7.82)

and substituting (7.80) into (7.66) leads to




s

s
s
+ 3s 2
s
,

z z z

(7.83)

if we define t = /(S03 )1/2 . A simple analytic solution is no longer possible, but


the development of the solution will be similar. The flux condition (7.82) at z = 0
allows the surface saturation to build up gradually, and a shock will only form if
1 (when the preceding solution becomes valid).

7.3.4 Similarity Solutions


If, on the other hand,  1, then the saturation profile approximately satisfies


s
s
=
s
,

z z
(7.84)

s
1 on z = 0,
s
=
0 on z = 1.
z
At least for small times, the model admits a similarity solution of the form
s = a f (),

= z/ b ,

(7.85)

7.3 Unsaturated Soils

405

Fig. 7.6 Schematic


representation of the
evolution of S for both large
and small

where satisfaction of the equations and boundary conditions requires 2a = b and


2b = 1 = a, whence a = 13 , b = 23 , and f satisfies
1
(ff  ) (f 2f  ) = 0,
3

(7.86)

with the condition at z = 0 becoming


ff  = 1

at = 0.

(7.87)

The condition at z = 1 can be satisfied for small enough , as we shall see, because Eq. (7.86) is degenerate, and f reaches zero in a finite distance, 0 , say, and
f = 0 for > 0 . As = 1/ 2/3 at z = 1, then this solution will satisfy the no flux
3/2
condition at z = 1 as long as < 0 , when the advancing front will reach z = 1.
To see why f behaves in this way, integrate once to find



2
f f  + = 1 +
f d.
(7.88)
3
0
For small , the right hand side is negative, and f is positive (to make physical
sense), so f decreases (and in fact f  < 23 ). For sufficiently small f (0) = f0 , f
will reach zero at a finite distance = 0 , and the solution must terminate. On the

other hand, for sufficiently large f0 , 0 f d reaches 1 at = 1 while f is still


2

positive (and f  = 23 1 there). For >
 1 , then f remains positive and f > 3
(f cannot reach zero for > 1 since 0 f d > 1 for > 1 ). Eventually f must
have a minimum and thereafter increase with . This is also unphysical, so we require f to reach zero at = 0 . This will occur for a range of f0 , and we have to
select f0 in order that
 0
f d = 1,
(7.89)
0

which in fact represents global conservation of mass. Figure 7.6 shows the schematic
form of solution both for 1 and  1. Evidently the solution for 1 will
have a profile with a travelling front between these two end cases.

406

7 Groundwater Flow

7.4 Immiscible Two-Phase Flows: The BuckleyLeverett


Equation
In some circumstances, the flow of more than one phase in a porous medium is
important. The type example is the flow of oil and gas, or oil and water (or all three!)
in a sedimentary basin, such as that beneath the North Sea. Suppose there are two
phases; denote the phases by subscripts 1 and 2, with fluid 2 being the wetting fluid,
and S is its saturation. Then the capillary suction characteristic is
p1 p2 = pc (S),

(7.90)

with the capillary suction pc being a positive, monotonically decreasing function of


saturation S; mass conservation takes the form
S
+ .u1 = 0,
t
S
+ .u2 = 0,

(7.91)

where is (constant) porosity, and Darcys law for each phase is


u1 =

k0

kr1 [p1 + 1 g k],


1

k0

u2 = kr2 [p2 + 2 g k],


2

(7.92)

with kri being the relative permeability of fluid i.


For example, if we consider a one-dimensional flow, with z pointing upwards,
then we can integrate (7.91) to yield the total flux
u1 + u2 = q(t).

(7.93)

If we define the mobilities of each fluid as


Mi =

k0
kri ,
i

(7.94)

then it is straightforward to derive the equation for S,





S
q
pc

+
=
+ (1 2 )g ,
Meff
t
z
M1
z

(7.95)

where the effective mobility is determined by



Meff =

1
1
+
M1 M2

1
.

(7.96)

7.4 Immiscible Two-Phase Flows: The BuckleyLeverett Equation

407

Fig. 7.7 Graph of


dimensionless wave speed
V (S) as a function of wetting
fluid saturation, indicating the
speed and direction of wave
motion (V > 0 means waves
move upwards) if the wetting
fluid is more dense. The
viscosity ratio r (see
(7.100)) is taken to be 30

This is a convective-diffusion equation for S. If suction is very small, we obtain


the BuckleyLeverett equation



S

q
+ (1 2 )g = 0,
(7.97)

+
Meff
t
z
M1
which is a nonlinear hyperbolic wave equation. As a typical situation, suppose
q = 0, and kr2 = S 3 , kr1 = (1 S)3 . Then
Meff =

k0 S 3 (1 S)3
,
1 S 3 + 2 (1 S)3

(7.98)

and the wave speed v(S) is given by



v = (2 1 )gMeff
(S) = v0 V (S),

(7.99)

where
v0 =

(2 1 )gk0
,
2

V (S) =

r
1
(S) =
+ ,
(1 S)3 S 3

 (S)
,
(S)2

1
r =
.
2

(7.100)

The variation of V with S is shown in Fig. 7.7. For 2 > 1 (as for oil and water,
where water is the wetting phase), waves move upwards at low water saturation and
downwards at high saturation.

Shocks will form, but these are smoothed by the diffusion term z
[Meff pc S
z ],
in which the diffusion coefficient is
D = Meff pc .

(7.101)

As a typical example, take


pc =

p0 (1 S)1
S 2

(7.102)

408

7 Groundwater Flow

with i > 0. Then we find



D = k0 p0 S

22

(1 S)

2+1


1 S + 2 (1 S)
,
1 S 3 + 2 (1 S)3

(7.103)

and we see that D is typically degenerate at S = 0. In particular, if 2 < 2, then


infiltration of the wetting phase into the non-wetting phase proceeds at a finite rate,
and this always occurs for infiltration of the non-wetting phase into the wetting
phase.
A particular limiting case is when one phase is much less dense than the other,
the usual situation being that of gas and liquid. This is exemplified by the problem
of snow-melt run-off considered earlier. In that case, water is the wetting phase, thus
2 1 = w a is positive, and also w 103 Pa s, a 105 Pa s, whence
a  w (r  1), so that, from (7.98),
Meff

k0 S 3
,
w

(7.104)

at least for saturations not close to unity. Shocks form and propagate downwards
(since 2 > 1 ). The presence of non-zero flux q < 0 does not affect this statement.
Interestingly, the approximation (7.104) will always break down at sufficiently high
saturation. Inspection of V (S) for r = 0.01 (as for air and water) indicates that
(7.104) is an excellent approximation for S  0.5, but not for S  0.6; for S  0.76,
V is positive and waves move upwards. As r 0, the right hand hump in Fig. 7.7
moves towards S = 1, but does not disappear; indeed the value of the maximum
1/3
increases, and is V r . Thus the single phase approximation for unsaturated
flow is a singular approximation when r  1 and 1 S  1.

7.5 Heterogeneous Porous Media


Perhaps the major concern in groundwater studies concerns the permeability.
Whereas we tend to think of the permeability as a well-defined quantity which reflects the local soil or rock properties, in reality it varies over many orders of magnitude on very small length scales. The consequence of this is that the value of the
permeability itself needs to be averaged in some way.
Permeability is so variable because of soil and rock heterogeneity. Because it
scales with the square of the constituent grain size, clay and sand permeabilities are
vastly different. And because sediments are lain down so slowly, over millions of
years, sand and clay layers often lie in close proximity. The same is true for sedimentary rocks, which are simply the same sand and clay layers cemented together
after burial and consequent subjection to high pressure and temperature.
In seeking to quantify porous medium flow at the large scale, we need to average the permeability in some way at the mesoscale: larger than the pore scale,
but less than the macroscale. The simplest approach is to suppose that the per-

7.5 Heterogeneous Porous Media

409

meability in a mesoscale block has a random distribution, often assumed to be


a lognormal distribution. An averaged permeability can be derived by supposing,
for example, that fluctuations have small amplitude. One finds that the consequent
averaged permeability is a tensor, whose components depend on the direction of
flow. In the following section, we consider a more specific model of the mesoscale
structure, where the heterogeneity is related to the occurrence of fractures in the
medium. This leads to the idea of a secondary porosity associated with the fractures.

7.5.1 Dual Porosity Models


Take a walk on exposed basement rock: at the seaside, in the mountains. Rocks are
not uniform, but are inevitably fractured, or jointed. There are numerous reasons for
this. Sedimentary rocks are lain down over millions of years via the deposition of
outwash clays, sands or calcareous microfossils in marine environments. Over this
time the deposition rate may average a millimetre or less per year. A metre of rock
may take a million, or ten million years, to accumulate. In this time, sea level may
rise or fall by tens or more of metres, and the land itself rises or falls because of
tectonic processes: the crashing of continents, the uplift of mountains, the burial of
sedimentary basins.
It is no surprise that in an exposed sedimentary sequence, such as one sees in
coastal cliffs, rocks form stratigraphic layers separated by unconformities marking
different sedimentary epochs. These unconformities are layers of weakness, and
when the rocks are later subjected to tectonic compression and folding, fractures
will form.
It is not only sedimentary rocks which tear as they are stressed. Igneous rocks
fracture as they solidify because of solidification shrinkage. They also form intrusions such as dikes and sills, whose different erosional properties can cause subsequent voidage.
The occurrence of faulting or jointing in rocks leads to a particular problem in the
description of groundwater flow through them. The rock itself is porous, and admits
a Darcy flow through its pore space; but the fractures act as a second porosity, admitting a secondary flow which would occur even if the rock itself was completely
impermeable. The situation is illustrated in Fig. 7.8. It is because of this configuration that the system is called a double, or dual, porosity system, and the resulting
model to describe the flow is called a dual porosity model.
In order to characterise porous flow through such a medium, we distinguish between the blocks of the matrix and the cross-cutting fractures. We suppose the fractures are tabular, or planar, of width h, and the blocks are of dimension dB , and
that h  dB . We denote the blocks by the domain M, and the fractures as M. Because the fractures are narrow, M essentially represents the external surfaces of
the blocks. We also suppose that dB  l, where l is a relevant macroscopic length

410

7 Groundwater Flow

Fig. 7.8 A doubly porous


system. Porous matrix blocks
are transected by (here) two
sets of transverse fractures

scale. For these tabular cracks, we can define a fracture porosity


f =

h
.
dB + h

(7.105)

We define a matrix pore pressure pm , which is the locally averaged pore pressure
in the matrix blocks, and a fracture pressure pf . There is then a matrix volume flux
per unit area um , and in the usual way we have Darcys law in the form
um =

km
pm ,

(7.106)

where km is the permeability of the fine-grained matrix. We have


km =

dp2
m

(7.107)

where dp is grain size and m is a tortuosity factor.


We suppose that flow in the fractures is essentially Poiseuille flow, and this leads
to a prescription of fracture volume flux per unit transverse width of crack, qf ,
(through a single fracture) as
qf =

h3
pf ,
f

(7.108)

where for a plane walled crack of width h, the fracture tortuosity f = 12; for rough
cracks, one can expect a higher value to be appropriate. The mean fracture velocity

7.5 Heterogeneous Porous Media

411

(which is also the fracture volume flux per unit area of fracture) is thus
uf =

kf
qf
= pf ,
h

(7.109)

where we define the fracture permeability parameter kf as


kf =

h2
.
f

(7.110)

Now if we consider the total (averaged) Darcy flux u through such a doubly porous
medium, it is straightforward to show that
u = (1 f )um M + f uf M ,

(7.111)

where the angle brackets denote averages: for um , a volume average over the matrix
blocks; for uf , an average over the fractures. Since h  dB , we can effectively
consider the average of uf to be a surface average over the fracture surface denoted
by M, the external boundary of the matrix blocks M. Note that each fracture has
two walls, and thus provides two external surfaces to M.
Our object is to characterise these averages in terms of macroscopic variables,
if possible. Notice that we have already carried out a primary averaging in defining
the fluxes um and uf in the first place: um is averaged over the grain scale of the
matrix, and uf is averaged over the width of the fractures. However, these fluxes
still represent values at a point within the larger block/fracture system. In particular,
note that by its definition the fracture flux is parallel to the fracture, and this carries
the implication that
pf
= 0,
(7.112)
n
where n is the normal to M (and we take it to point from the matrix into the
fracture).
We now want to average over the larger block scale dB . The point fluxes um
and uf satisfy the conservation of mass equations
.um = 0

(7.113)

.(huf ) = um .n|M ,

(7.114)

and

where in (7.114) there is a flux um .n at the fracture surfaces from the matrix to the
fractures. Some comment on this equation is necessary. It takes this form because
of the fact that the fracture flux as defined in (7.108) is already averaged over the
cross section of the fracture. Continuity of the fluxes at the block fracture interface
produces the source term in (7.114) through the integration of the fracture point
transverse velocity across the fracture.

412

7 Groundwater Flow

To use the ideas of homogenisation, we first define dimensionless variables. We


scale the variables as
uk U,

pk P ,

x l,

(7.115)

where l is the macroscopic length scale, and we define a second (now dimensionless)
spatial variable X by putting
x = X,

(7.116)

where
dB
.
(7.117)
l
The blocks thus have size X O(1). We write, with an obvious notation, the dimensionless gradient operator in the form
=

1
= x + X ,

(7.118)

where we are now using x and X as multiple spatial scales. We suppose that the
block structure is periodic in X, although this is inessential for the methodology.
p
Now the requirement of (7.112) that nf = 0 implies


1
n. x pf + X pf = 0,
(7.119)

whence it follows that we can write, approximately,


pf = p(x) + p f (X),

(7.120)

and then
p f
p
n. x p = n. X p f
.
(7.121)
n
N
p is the macroscopic average pressure variable, and we may impose periodicity in
X of p f with zero mean.
We have continuity of matrix and fracture pressure at M, and therefore we can
write
pm = p(x) + p m (X),

(7.122)

and the matrix pressure satisfies


X2 p m = 0 in M,
p m = p f

(7.123)

on M.

To find the solution pm of (7.123), define a Greens function G(X, Y) which satisfies
Y2 G = (X Y) in M,

G = 0 for Y M.

(7.124)

7.5 Heterogeneous Porous Media

Then p m is given by

413


p m =
M

where

NY

G(X, Y)
p f (Y) dS(Y),
NY

(7.125)

= n(Y). Y , and it follows from this that (for X M)


p m
=
N


K(X, Y) p f (Y) dS(Y),

(7.126)

2 G(X, Y)
.
NX NY

(7.127)

where
K(X, Y) =

It remains to determine the fracture pressure perturbation p f . This involves solving (7.109) and (7.114). Supposing h and kf are constant, these reduce, at leading
order in , to



p m p
dB km
2
X p f =
+
,
(7.128)
hkf
N
n
subject to conditions of periodicity in X and zero mean. Note that the Laplacian in
(7.128) is defined on the surface M. Using (7.126), we can write (7.128) in the
form


p
K(X, Y) p f (Y) dS(Y) +
,
(7.129)
X2 p f =
n
M
where
=

dB km f dB dp2
=
.
hkf
m h3

The canonical microscale fracture problem to be solved on M is thus




2
K(X, Y) q(Y) dS(Y) + n ,
M q =

(7.130)

(7.131)

with periodic boundary conditions and zero mean, and then


p f = q(X). x p.

(7.132)

We now use these results to find the effective permeability of the medium. Averaging (7.106) over the matrix blocks yields (dimensionlessly)
U um M =

km P
x p,
l

(7.133)

since pm is continuous across fractures and periodic in X. In a similar way,


U uf M =


kf P

x p +  X p f M ,
l

(7.134)

414

7 Groundwater Flow

but the surface average term in this expression does not obviously vanish. We have


X p f = X p f n(n.X p f ) + n(n. X p f );
(7.135)
the term in square brackets is a tangential derivative of p f along M, and we separate the terms in this way because pf is only defined on M. In addition, because
p f = p m on M, we could replace the subscript f by m in the square-bracketed
expression. Because of (7.121), we have
n(n.X p f ) = n(n. x p),

(7.136)


p
,
n(n.X p f ) M = ei ni nj M
xj

(7.137)

and thus


where ei is the unit vector in the xi direction. We therefore have


U uf M =




kf P

I nn M . x p + X p f n(n.X pf ) M ,
l

(7.138)

where I is the unit tensor, and nn is the tensor with elements ni nj .


We now substitute the expression for p f in (7.131) into (7.138); averaging over
M, we finally derive the expression for the mean fracture velocity in the dimensionless form
Uuf M =


 
kf P

I nn M + (I nn). X q . x p.
l

(7.139)

Rewriting this in dimensional form, we have


uf M =

kf
k .p,

where the fracture relative permeability tensor k is defined by





qj

.
kij = (ik ni nk ) j k +
Xk M

(7.140)

(7.141)

Equally, the dimensional matrix flux is, from (7.133),


um M =

km
p.

(7.142)

(7.140), (7.141) and (7.142) give the recipes for the averaged matrix and fracture
fluxes in terms of the macroscopic pressure gradient and the solution of the block
scale fracture pressure problem (7.131).
If we take a representative volume consisting of many blocks, and integrate
(7.113) over the matrix volume, and (7.114) over the fracture volume, we obtain

7.6 Contaminant Transport

415

the averaged (dimensional) equations for the averaged fluxes in the form

. f uf M = sf um .n M ,

. (1 f )um M = sf um .n M ,

(7.143)

where sf is the specific fracture surface area (i.e., surface area per unit volume: here
sf dB1 ). Note that the source term in (7.143) is just


km sf P pm
sf um .n M =
,
(7.144)
l
N M
6
because  p
n M = n M . x p = 0, and in the present case this is just zero because
of (7.123).
The usefulness of all this methodology is that it carries across to other, more
complicated averaging problems (as we see below), but in the present case of incompressible double porosity flow, it may be somewhat unnecessary. The reason
for this is that fracture relative permeability depends on the solution of (7.131), and
thus on the fracture geometry and the single dimensionless parameter given by
(7.130). Assuming small fracture porosity, the ratio of matrix flow to fracture flow
is, from (7.111), (7.109), (7.105) and (7.106), of the order of

km dB
um

= .
uf
kf h

(7.145)

If is large, then very little flow occurs through the fractures anyway, and the secondary porosity is of little concern. If is small, the blocks are essentially impermeable, and the fracture network is crucial; but then the solution of (7.131) is just
q = O(), and the relative permeability is simply
k = I nn M ,

(7.146)

which only differs from the unit tensor if the medium is anisotropic. It is only in the
case = O(1) that the competition between the two systems becomes important. If
we use values f = 102 , dB = 1 m, h = 103 m and km = 1012 m2 (cf. Table 7.1),
we get 0.01. This might be appropriate for a fractured sandstone on a regional
scale. Generally, the primary and secondary (fracture) permeability will only be
comparable if the host rock is itself quite permeable.

7.6 Contaminant Transport


Much of the interest in modelling groundwater flow lies in the prediction of solute
transport, in particular in understanding how pollutants will disperse: for example,
6 The

average of n over M is zero because


the normals
on opposite


 sides of a fracture are in
opposite directions. More specifically, M n dS = M dV , thus M n dS = 0.

416

7 Groundwater Flow

how do nitrates used for agricultural purposes disperse via the local groundwater
system? Mostly simply, one would simply add a diffusion term to the advection of
the solute concentration c:
ct + u.c = .[Dc].

(7.147)

The diffusive width l of a sharp front travelling at speed u after it has travelled a distance l is of the order of l (Dl/u)1/2 ; if we take D 109 m2 s1 ,
u 106 m s1 (30 m y1 ), l = 103 m, then l 1 m, and the diffusion zone is
relatively narrow. For a more porous sand, the diffusion width is even smaller.
In fact, as velocity increases, the effect of diffusion increases. That this is so is
due to a remarkable phenomenon called Taylor dispersion, described by G.I. Taylor
in 1953. Consider the diffusion of a solute in a tube of circular cross section through
which a Poiseuille flow passes. If the mean velocity is U and the tube is of radius a,
then the velocity is 2U (1 r 2 /a 2 ), and the concentration satisfies the equation




1
2 2
ct + 2U 1 r /a cx = D crr + cr + cxx ,
(7.148)
r
where x is measured along the tube, and r is the radial coordinate. Taylor showed,
rather ingenuously, that when the Pclet number Pe = aU/D is large, then the effect
of the diffusion term in (7.148) is to disperse the mean solute concentration diffusively about the position of its centre of mass, x = U t, with a dispersion coefficient
of a 2 U 2 /48D. Aris later improved this to
DT =

a2U 2
+ D,
48D

(7.149)

which is asymptotically valid for x a. The dispersive mechanism is due to the


radial variation of the velocity profile, which can disperse the solute even if the
diffusion coefficient is very small.
Typically, this is generalised for porous media (where we think of the pores as
being like Taylors tube) by writing the dispersion coefficient as
D T = D + D ,

(7.150)

where D represents molecular diffusion and D dispersion in the direction of flow.


The tortuosity of the flow paths and the possibility of adsorption on to the solid
causes D to be less than D, and ratios D /D between 0.01 and 0.5 are commonly
observed. In porous media, remixing at pore junctions causes the dependence of D
on the flow velocity to be less than quadratic, and a relation of the form
D = um ,

(7.151)

where u is the Darcy flux, fits experimental data reasonably well for values 1 <
m < 1.2. A common assumption is to take m = 1. Mixing at junctions also causes
transverse dispersion to occur, with a coefficient D which is measured to be less
than D by a factor of order 102 when Pe 1. Dispersion is thus a tensor property.

7.6 Contaminant Transport

417

If we write
D =  |u|

(7.152)

for the longitudinal dispersion coefficient, and


D = |u|

(7.153)

for the lateral dispersion coefficient, then a suitable tensor generalisation is


DijT = |u|ij + ( )

ui uj
,
|u|

(7.154)

where ij is the Kronecker delta.


The conservation of solute equation is then


T

c
T c
Dij
.
+ u.c = . D .c =
t
xi
xj
For a one-dimensional flow in the x direction, c satisfies






c

c
c
c
c
+v
=
D
+
D
+
D
t
x x
x
y
y
z
z

(7.155)

(7.156)

(v = u/ is the linear velocity) and if the dispersivities are constant, then the solution for release of a mass M at the origin at t = 0 is


M
(x vt)2
r2
c=
exp

,
4D t
4D t
8(D )1/2 D t 3/2

(7.157)

where r 2 = y 2 + z2 .

7.6.1 Reactive Dual Porosity Models


Let us now consider the reactive transport of a contaminant of concentration c within
a fractured soil or rock which has dual porosity. We follow the ideas of averaging
and homogenisation in Sect. 7.5.1. The point forms of the equations are taken in the
form
c
+ .(cu) = .[Dc] + Sf ,
t
cm
+ .(cm u) = .[Dm .cm ] + Sm ,
t

(7.158)

where Sf , Sm represent source or sink terms due to chemical reaction, D is the


molecular diffusion coefficient of the contaminant within the fractures, and Dm is

418

7 Groundwater Flow

the dispersivity within the matrix blocks. The concentrations within the fractures
and matrix are denoted by c and cm , respectively.
The first thing to do is to average the fracture concentration equation across the
width of the fracture. When we do this, we effectively regain the Taylor dispersion
equation, with the addition of the reaction terms, and also a solute flux delivered
from the matrix:

cf
1

+ .(cf uf ) = .[Df .cf ] + Sf (n.Dm .cm ) cm um .n M . (7.159)


t
h
This equation is analogous to (7.114). It differs from (7.158) in that cf is the crosssectional average concentration (actually, in the derivation of the Taylor dispersion
equation, one finds the concentration is cross-sectionally uniform, so that cf = c);
uf is the cross-sectionally averaged fracture velocity, just as before; and Df is the
local dispersion coefficient: it will be modified again at the larger macroscale. The
reaction term Sf depends on cross-sectionally averaged concentrations, which equal
their point forms, so that Sf is unchanged.
Now we write down the equivalents of (7.143). We define the block averaged
concentrations
cm = cm |M ,

cf = cf |M ,

(7.160)

where as before c |M denotes an average of c over the matrix blocks M, and c |M


denotes the average of c over the fracture surface M. Then we have the equation
for cf :

(f cf ) + .[f cf uf ] = .[f Df . cf ] + f Sf
t


+ sf n.{cm um Dm .cm } M .

(7.161)

We can make use of (7.143)1 , and the fact that cf = cm on M, to simplify this to

f


cf
+ uf . cf = .[f DF . cf ] + f Sf sf n.Dm .cm |M . (7.162)
t

The specific fracture surface area is defined as sf , as before. The macroscopic fracture dispersivity DF here is distinct from Df , in the same way that Taylor dispersion
in a tube is distinct from that in a porous medium; in this case it is because of remixing of fracture fluid at the junctions between fractures at the block boundaries. The
formal averaging assumption which is made is
Df .cf cf uf |M = DF . cf cf uf |M ,

(7.163)

and this has some justification insofar as it is just this result which emerges in the
study of Taylor dispersion.

7.6 Contaminant Transport

419

In a similar way to the derivation of the average fracture concentration equation,


the matrix averaged concentration satisfies


cm
(1 f )
+ um . cm
t


= . (1 f )Dm . cm + (1 f )Sm + sf n.Dm .cm |M .
(7.164)
The result of averaging is the two Eqs. (7.162) and (7.164) for the average
fracture and matrix concentrations. In principle, the block average fracture dispersivity DF should be calculable by solving the local block problem, although
in practice one would assume a value by analogy with assumptions about porous
medium dispersion coefficients. However, unlike the incompressible dual porosity
mass flow equations (7.143), the source term sf n.Dm .cm |M is non-zero, and
this must be constituted, ideally by solving the block scale problem, which is given
by Eqs. (7.158)2 and (7.159); these can be slightly simplified to the forms
cm
+ um .cm = .[Dm .cm ] + Sm ,
t
cf
1
+ uf .cf = .[Df .cf ] + Sf (n.Dm .cm )|M .
t
h

(7.165)

The boundary condition for cm is that


cm = cf

on M,

(7.166)

and cf is periodic over M.


The basis for the method of homogenisation is the expansion of the local block
problem in terms of the parameter = dB / l, where dB is the block scale and l is
the macroscopic length scale. However, because of the complexity of the equations
to be solved, the application of this method must be done judiciously.
To illustrate this point, suppose that the dispersion coefficient tensors are all
isotropic, and equal to DT I, where DT is constant, and that suitable (macroscopic)
scales for the variables are c c0 , x l, u U , t l/U , and suppose to be precise
that the reaction kinetics are first order, i.e., S = rc, and that the specific fracture surface area is constant, sf = 1/dB . The dimensionless equation for the matrix
average cm is then


cm
+ um .x cm
(1 f )Pe
t


1 cm 
= (1 f )x2 cm Pe(1 f )cm + 2
,
(7.167)
N M
where

cm
N

denotes n. X cm , and
Pe =

Ul
,
DT

rl
.
U

(7.168)

420

7 Groundwater Flow

The local block equation (7.165)1 rewritten in the block variables


x
X= ,

is then

t
T= ,



cm
Pe
+ um .X cm = X2 cm 2 Pecm .
T

(7.169)

(7.170)

What is obvious is that all the terms cannot balance in both local and global problems, and this leads to simplifications. The simplest case is where the macroscopic
Pclet number and reaction number are both taken to be O(1); then (7.167) suggests that in (7.170) we put
+ 2 cm (X)
cm = c(x)

(7.171)

+ 2 cf (X)); both inner and outer problems are well scaled,


(and thus also cf = c(x)
and the advective term can be neglected in solving the inner problem, which in fact
reduces approximately to Poissons equation. Specifically, using multiple scales in
both x and t, we have at leading order in ,


X2 cm


c
+ um . x c + Pec x2 c
= Pe
t

(7.172)

subject to cm = cf on M and conditions of periodicity on M. (If we integrate


(7.172) over M, we regain the averaged equation (7.164) as an integrability condition for (7.172), as expected via homogenisation.)
In a similar way, the leading order local equation for the fracture perturbed concentration cf is
X2 cf




(1 f ) cm 
c
+
u

=
Pe
.
c

+ Pec x2 c
f
x
f
N M
t

(7.173)

c
|M x c.n|
M is identically zero, because the normals
(note that a term ff n
n on the two faces of a fracture cancel each other). cf satisfying (7.173) is subject to
periodicity in M and zero mean. As for the fluid flow, we can solve (7.172) using
a Greens function for the Laplacian in M, so that the boundary derivative term
in (7.173) becomes an integral convolution in terms of cf . We then solve (7.173)
using a Greens function for the Laplacian on M. This allows us to determine the
homogenised equation for c (see Question 7.13).
In fact, it is rarely the case that Pe and are O(1): more commonly they are both
large. Suppose for example that = 104 , Pe = 104 , = 104 : not unreasonable
values (as we shall see below). Putting
(1 )

p = Pe,

= ,

(7.174)

7.7 Environmental Remediation

421

it seems natural to take p O(1). The block problem becomes (in terms of T
and X only)


cm
+ um . X cm = X2 cm pcm ,
p
(7.175)
T
which implies that reaction occurs on the block scale. This is a linear equation for cm
which can in principle be solved to give the boundary flux term in the cf equation
as a convolution integral in terms of cf . In a strict sense, this distinguished limit
describes the structure of a reaction front in which both reaction and dispersion are
important. Outside this front, reaction and dispersion are negligible, and the reactant
simply advects with the flow.7 Importantly, it implies that the speed of the front is
determined by the local diffusion and reaction within the block.
If is even larger than this, so that 1, then the reaction is fast at the block
scale, and occurs in a thin rind within the block; for a single species, as here, this
rind must be on the boundary, as the interior reactant concentration reaches zero
rapidly. Bearing in mind that the normal coordinate n at M points into the fracture,
the boundary layer solution for cm is just


(7.176)
cm cf exp (p)1/2 n ,
and thus the flux term in the local fracture equation derived from (7.165)2 is

1
1 cm 
= (p)1/2 cf .
(7.177)

f n M
f
Hence cf satisfies the local equation
X2 cf



cf
1
1/2

(p) cf = p
+ uf .X cf + pcf .
f
T

(7.178)

As is perhaps obvious, the reaction is fast in the fractures also and cf rapidly approaches zero.

7.7 Environmental Remediation


Environmental pollution is now an endemic phenomenon. All over every industrialised country, spills of hydrocarbons, effluents, and industrial waste have caused
pollution of underlying groundwater. The resulting plumes move slowly with the
prevailing background groundwater flow, and in course of time will enter streams,
rivers and lakes, with consequent health risk via the contamination of drinking water. Short of more drastic measures such as pumping out polluted groundwater and
this only makes sense for multi-species reactions of the form A + B P , for then the
reaction rate AB, and can be zero by virtue of A = 0 on one side of the front, and B = 0 on the
other.

7 Actually,

422

7 Groundwater Flow

treating it, natural bioremediation anticipates that microbial action will eventually
break down most pollutants, rendering them harmless. The issue for the environmental scientist is to predict the future movement of the plume, and the likelihood
of microbial breakdown before it reaches drinking water sources.
In so doing, groundwater flow modelling is essential, because of its ability to
predict into the future, and also because accurate monitoring of subsurface contamination is expensive and not straightforward. Against this, subsurface soil and rock
is usually an extremely heterogeneous medium, both physically and chemically, and
the validation of computational results is difficult.

7.7.1 Reactive Groundwater Flow


The general context of many subsurface pollution problems of concern is similar,
and we therefore begin with some generalities. Contaminants may be aqueous, in
which case they mix with the groundwater, or non-aqueous, in which case they
do not. Hydrocarbons, for example, are non-aqueous. Amongst the non-aqueous
phase liquids (NAPLs), one distinguishes dense liquids (DNAPLs) from light ones
(LNAPLs). DNAPLs will sink into the saturated zone, whereas LNAPLs, such as
hydrocarbons, will sink to the base of the unsaturated zone, and there sit on the
water table, from where their constituents may diffuse downwards.
The contaminant plume will typically consist of a cocktail of different chemicals,
which flow with the local groundwater flow, disperse within it, and react with oxygen and other substances in the soil via the agency of microbial action. The typical
sort of model of concern is thus the reaction-advection-dispersion equation
R

c
+ .(cu) = .(D .c) + S,
t

(7.179)

where c is one of a sequence of reactants, u is the local groundwater flux (given by


Darcys law), D represents dispersion, and is typically anisotropic, in the sense that
dispersion in the longitudinal direction is larger than lateral dispersion. Dispersion
itself is partly due to molecular diffusion, but more importantly (at high grain scale
Pclet number) is due to grain scale shear-induced distortion of the fluid associated
with Taylor dispersion, together with remixing at pore junctions. Typically, the longitudinal dispersion coefficient D dp |u|, where dp is grain size, while lateral
dispersion D is a factor of 10100 smaller.
The coefficient R is the retardation factor, and it is a slowing rate due to the adsorption of aqueous phase concentration on solid particles. Specifically, we actually
have two separate conservation laws for solid and aqueous concentrations cs and cl ,
respectively, thus
cl
+ .(cl u) = .(D.cl ) + S + kd cs ka cl ,
t
cs
= kd cs + ka cl ,
t

(7.180)

7.7 Environmental Remediation

423

where kd and ka are desorption and adsorption rates, respectively. (7.180) specifically assumes that the reaction occurs only in the aqueous phase, by way of example.
Now the point is that if the sorption rates are very large (and constant), then the
second equation in (7.180) tells us that
cs

ka cl
,
kd

(7.181)

and the sum of the two equations then gives us (7.179), with the retardation factor
R=1+

ka
.
kd

(7.182)

The source term S represents reaction driven sources and sinks. It is typically
the case that there are many, many reactions and reactants. Equally typically, many
of the reaction rates are not well known, and the rates may be widely disparate. In
general this will imply that many reactions can be taken to be in equilibrium, with
only the slowest (rate-controlling) reaction being of dynamical importance.

7.7.2 Biomass Modelling


The reaction terms S in (7.180) are mediated by bacteria within soil, which consume
the various nutrients provided through metabolic reactions. Like all living things,
microbes survive by generating energy from nutrients through a variety of such reactions. This process involves a network of redox (oxidationreduction) reactions,
and involves the overall exchange of electrons between two distinct chemical fuels
which are consumed in the reactions; the metabolic process is in this case called
respiration. While there may be a number of such fuels, there is a hierarchy in their
use. Dissolved oxygen is commonly the terminal electron acceptor (as the externally sourced oxidant is typically referred to), while an organic carbon compound
is the electron donor. When these preferred substrates are absent or depleted, other
compounds can be used instead. When the same organic compound is used as both
donor and acceptor, the metabolic process is called fermentation.
Many bacteria are able to use several reaction pathways independently, giving
them a degree of flexibility to differing conditions. This capability is very speciesdependent, and competition ensures that the species which are best adapted to
local conditions become dominant. Microbial growth depends heavily on energy
metabolism but also requires the uptake of other substrates needed to generate new
biomass. Growth rate is generally limited by the supply of one or more substrates,
but saturates to a maximum growth rate in conditions of ample supply. The dependence of bacterial growth rate is commonly taken, by analogy with simple enzyme
c
kinetic uptake rates, as proportional to K+c
, where c is the relevant nutrient concentration, and K is a constant; such kinetics are called Monod kinetics. When two

424

7 Groundwater Flow

nutrients control growth, as in respiration, it is usual to take the growth rate as proportional to the product of two Monod factors, thus (for example)
S = r0 X

c2
c1
,
K1 + c1 K2 + c2

(7.183)

in which r0 is a reaction rate constant, and X is the biomass density.


More complex models also consider the growth of the microbial population X
in terms of the nutrient supply, in particular in the form of a microbial mat which
becomes attached to the soil grains, and usually called a biofilm. In particular, the
concept of biodegradation, for example of oil spills in sea water, or of factory effluent in groundwater, commonly involves the development of bacterial colonies which
are able to efficiently use the offending contaminant to promote their own growth.

7.7.3 Non-dimensionalisation
It is commonly the case that the water table is at a depth of 1020 metres, whereas a
plume may have spread (or we are concerned with whether it will spread) a distance
of order kilometres. Therefore these plumes generally have high aspect ratio, a cause
both of computational stress and analytical simplification. The latter may be offset
by the decreased lateral dispersion coefficient, as we now show.
Let us consider the scalar contaminant equation (7.179) in two dimensions, with
horizontal and vertical coordinates x and z, corresponding Darcy fluxes u and w,
and longitudinal (horizontal) and transverse (vertical) dispersivities D and D ,
which we take to be constant. We suppose l and d are suitable horizontal and vertical
length scales, U is a horizontal Darcy flux scale (and therefore mass conservation
implies that hU/ l is a suitable vertical flux scale), Rl/U is then the convective time
scale, and we take S0 to be a measure of the reaction rate term, and c0 to be a typical
contaminant concentration. The units of concentration are mol l1 (moles per litre),
and the units of S0 are mol l1 s1 .
We define dimensionless variables by writing
u = U u ,
x = lx ,

w=

hU
w ,
d

z = dz ,

c = c0 c ,

t=

Rl
t ,
U

(7.184)

S = S0 S;

substituting these into (7.179) (and forthwith dropping the asterisks), we obtain the
dimensionless equation
c
1 2c
c
c
1 2c
+
+ S,
+u
+w =
t
x
z Pe x 2 Pe z2

(7.185)

where
Pe =

Ul
,
D

Pe =

U d2
,
D l

S0 l
.
U c0

(7.186)

7.8 Three Specific Remediation Problems

425

The aspect ratio d/ l and transverse dispersivity ratio D /D compete against
each other, and generally we might suppose Pe Pe ; but also if D dp U , then
we can expect that usually Pe 1, heralding the existence of thin boundary layers
in which dispersion is effective. The parameter is the ratio of advection time
to reaction time, and will often be very large for microbially mediated reactions
of interest. These two observations cause numerical difficulties in solving (7.186),
but can aid analytical insight. In the following section, we discuss three specific
groundwater contamination problems of recent concern.

7.8 Three Specific Remediation Problems


Four Ashes Four Ashes is a site in the Midlands of England where a fifty year
old plume of phenol (C6 H6 0) and other contaminants, about 500 metres long, lies
within the saturated zone. The plume is thought to lie between depths of 10 m and
30 m at 130 m from the source, but is sinking as it moves west (perhaps because
of surface recharge), reaching depths 2144 m at 350 m distance (this information
comes from two boreholes drilled at the site). Degradation is very slow: perhaps
only 5% of the total contaminant load has so far been degraded. This is thought
to be due to toxicity of the phenol within the plume, and to supply limitation of
electron acceptors (oxygen and nitrate) at the plume fringe.
The overall reaction scheme which is considered to apply is that the phenol is
oxidised at the plume boundary by oxygenated groundwater, producing TIC (total inorganic carbon). Within the plume, phenol is fermented to produce hydrogen
amongst other things, which then reduces various oxides within the soil via microbial agency. A suitable set of reactions to describe the situation consists of the
following:
r1

+
C6 H6 O + 7O2 + 3H2 O 6HCO
3 + 6H ,
r2

+
C6 H6 O + 5.6NO
3 + 0.2H2 O 6HCO3 + 0.4H + 2.8N2 ,
r3

C6 H6 O + 5H2 O 3CH3 COOH + 2H2 ,


r4

+
C6 H6 O + 17H2 O 6HCO
3 + 6H + 14H2 ,
r5

H2 + MnO2(s) + 2H+ 2H2 O + Mn2+ ,

(7.187)

r6

H2 + 2FeOOH(s) + 4H+ 4H2 O + 2Fe2+ ,


r7

H2 + SO2
4 + 0.25H H2 O + 0.25HS ,
r8

+
H2 + CO2
3 + 0.5H 0.75H2 O + 0.25CH4 ,

and r1 r8 are the reaction rates, defined using Monod kinetics. In terms of the reaction rates, the source term in each reactant equation is then

Sj =
sj l rl ,
(7.188)
l

426

7 Groundwater Flow

where sj l is the (sign dependent) stoichiometric coefficient for reactant j in reaction l.


The maximum (saturated) rates kj of the Monod rates rj are thought to range
from about 1013 mol l1 s1 to about 1010 mol l1 s1 . The background phenol
concentration p0 is of the order of 1010 mol l1 , and the aqueous dissolved oxygen
level c0 is of order 104 mol l1 . The horizontal Darcy flux is estimated to be U =
10 m y1 . The dispersivities are taken in the form D = U , and values of  = 1 m
and = 4 104 m are assumed.8 Also, we take l = 500 m, d = 20 m, and R = 1.
With all of these values, we have
Pe =

l
= 500,


d2
= 2000,
l

Pe =

106 109 ,

(7.189)

where the definition of is based on p0 , i.e.,


=

S0 l
.
Up0

(7.190)

As suggested, Pe Pe 1, and the reaction rate parameter is extremely large.


To get an idea of the solution behaviour when Pe 1 and 1, we consider the
simpler reaction
r

C6 H6 O + O2 products,

(7.191)

where is a suitable stoichiometric coefficient. The corresponding version of


(7.185) is simply9
p
+ .(pu) =
t
c
+ .(cu) =
t

1 2
p + S,
Pe
1 2
c + S,
Pe

(7.192)

where p and c are the concentrations of the two species (phenol and oxygen), and
the reaction term is just



c
p
;
(7.193)
S=
kp + p
kc + c
the parameter is given by
=

p0
106 .
c0

(7.194)

both theoretical and experimental reasons, we expect  dp , the pore or grain size (Bear
1972, p. 609), with / 0.010.05 (Sahimi 1995, p. 225). The large value of  here suggests
flow in sub-parallel fractures with spacing on the order of a metre.

8 For

9 We

can take Pe = Pe by choosing d appropriately.

7.8 Three Specific Remediation Problems

427

The asymptotic structure of the solution of (7.192) is easily given. For 1, the
reaction region is very thin, and will occur only at the fringe of the plume. Outside
the plume, p = 0, while inside the plume c = 0, so that S = 0 everywhere apart
from the reaction front, which is located at the plume boundary. Denoting the plume
by P and its boundary by P , we thus have to solve
p
+ .(pu) =
t
c
+ .(cu) =
t

1 2
p, x P ,
Pe
1 2
/ P,
c, x
Pe

(7.195)

and p and c both satisfy p = c = 0 on P . The location of P has to be determined


(it is a free boundary), but this is simply done because of the precise conservation
law for c p, which follows from (7.192):


(c p)
1
+ . (c p)u = 2 (c p).
t
Pe

(7.196)

Integrating this across the reaction front yields the extra condition10 to determine
the front P :
c
p
= .
(7.197)
n
n
Note that the normal derivative here has to be treated with some attention when
non-dimensionalised. If the plume fringe is at dimensionless position z = (x, t),
then the dimensional normal component of the dispersive flux of oxygen (in terms
of dimensionless variables) is just


1 c
1 c
c0 U d

,
(7.198)
(D.c).n =
l(1 + 2 x2 )1/2 Pe z Pe x x
where is the aspect ratio. There is a similar expression for p/n. It follows from
this that more generally, the conservation boundary condition (7.197) can be expressed as


1 p
1 c
1 p
1 c

,
(7.199)
Pe z Pe x x
Pe z Pe x x
and to interpret this as (7.197) requires the unit normal n to be suitably defined.
The problem is even easier when (as here) Pe 1. Then diffusive boundary layers of (dimensionless) thickness O(1/Pe1/2 ) adjoin the plume boundary, and away
from these, the reactants are simply advected with the flow.
In the present case, the parameter  1. This modifies the discussion as follows.
Reaction must still occur in thin regions, outside which the reactants obey (7.195),
10 In

this particular case, it is even simpler, since we just integrate the convective-diffusion equation (7.196) with appropriate external boundary conditions, and P is located by the curve where
c p = 0.

428

7 Groundwater Flow

but there cannot be a front (in the sense of a moving boundary) at which c =
If, for example, we have the one-dimensional problem
ct + ucx =

1
cxx ,
Pe

c
n

= 0.

(7.200)

with c = 1 at t = 0 and as x , and c = 0 at x = 0, then if Pe O(1), reaction


occurs near x = 0, oxygen supply being sufficient to remove the phenol. If Pe 1,
then a front moves away from x = 0 at speed u. Behind this front, c 1, ahead of
it c 0, and the front itself is purely diffusional, of width O{(t/Pe)1/2 }, having an
error function profile for solution.
Rexco The site was a former coal carbonisation plant (between 1935 and 1970)
where ammonium liquor had been allowed to drain on site. The liquid drained
through the unsaturated zone of some 18 metres depth to the saturated zone, where
it is spreading as a plume of some 25 metres depth in a sandstone aquifer. At the
plume fringe, ammonium ions (NH+
4 ) ions are subject to oxidation, releasing nitrogen. Phenol is also present in front of the ammonium, which is highly retarded (with
a retardation factor of R = 5). Well extraction by a new factory on site is causing
the groundwater flow field to be time dependent, and this is an issue in modelling
studies.
The modelling scheme is similar to the Four Ashes site, but less is known of the
reaction rates. The reactions which are considered to be important for the ammonium are
r1

+
NH+
4 + 2O2 NO3 + H2 O + 2H ,
r2

+
NO
3 + 1.25CH2 O + H 0.5N2 + 1.25CO2 + 1.75H2 O,

(7.201)

r3

+
NH+
4 + 0.6NO3 0.8N2 + 1.8H2 O + 0.4H .

The kinetics (only) of these reactions is described by the five ordinary differ
ential equations for the five reactants, ammonium NH+
4 , oxygen O2 , nitrate NO3 ,
+
hydrogen ion H and organic carbon CH2 O:
dcNH+

4
= r1 r3 ,
dt
dcO2
= 2r1 ,
dt
dcNO
3
= r1 r2 0.6r3 ,
dt
dcH+
= 2r1 r2 + 0.4r3 ,
dt
dcCH2 O
= 1.25r2 .
dt

(7.202)

7.8 Three Specific Remediation Problems

429

To estimate Pclet numbers, we take d = 25 m, l = 1000 m, R = 5, U =


100 m y1 . Values of dispersion coefficients are uncertain, so we choose for illustration D = 100 m2 y1 , D = 10 m2 y1 , i.e.,  = 1 m (like Four Ashes)
and = 0.1 m. The (unretarded) time scale l/U is then 10 years, and Pe 103 ,
Pe 6. These values are not too reliable, and perhaps are consistent with a highly
fractured aquifer. For a homogeneous medium, we would expect much smaller ,
and thus much larger Pe.
It is possible to extend the Four Ashes discussion of the simple phenol/oxygen
reaction to the more complicated scheme above. Reaction rates are not well known.
Most of the reactants are present at the level of mmol l1 (millimole per litre). Specific estimates are
cNH+ 12 mmol l1 ,
4

cH+ 104 mmol l1 ,

cO2 0.3 mmol l1 ,


cCH2 O 1 mmol l1 .

cNO 0.15 mmol l1 ,


3
(7.203)

If we use cNH+ = 12 mmol l1 as a concentration scale in the definition of in


4

(7.186) and arbitrarily pick a reaction rate scale of S0 = 1010 mol l1 s1 (comparable to phenol degradation rate at Four Ashes), then we have O(1), and the
simple description for Four Ashes is inappropriate.
Despite this, it is thought that reactions are fast, and we will suppose that in fact
1. In particular, there are data from a borehole measurement which appear
to be consistent with the idea that there is a thin reaction zone at the upper plume
boundary. Oxygen (presumably) diffuses to this front from above, and ammonium
from below; in the reaction zone there is a huge spike of nitrate (see Fig. 7.9). We
wish to see whether the existence of this nitrate spike is consistent with the model,
assuming the measurement was realistic. (Figure 7.9 actually suggests the existence
of two fronts, with nitrate produced at the upper front diffusing down to the second
front, but we will focus only on the upper front.)
The dimensionless equations equivalent to (7.192) for the reaction scheme
(7.201) are
2
2
1 cNH+4
1 cNH+4
+ .(c
u) =
+
(r1 + r3 ),
R
t
Pe z2
Pe x 2
1 2 cO2
1 2 cO2
cO2
+
2O2 r1 ,
+ .(cO2 u) =
t
Pe z2
Pe x 2
2
2
cNO
1 cNO3
1 cNO3
3
+
+ .(cNO u) =
3
t
Pe z2
Pe x 2
+ NO (r1 r2 0.6r3 ),

cNH+
4

NH+
4

(7.204)

cH+
1 2 cH+
1 2 cH+
+
+ H+ (2r1 r2 + 0.4r3 ),
+ .(cH+ u) =
t
Pe z2
Pe x 2
cCH2 O
1 2 cCH2 O
1 2 cCH2 O
+ .(cCH2 O u) =
+
1.25CH2 O r2 ,
2
t
Pe z
Pe x 2

430

7 Groundwater Flow

Fig. 7.9 Data from borehole 102 at Rexco, May 2003, courtesy of David Lerner and Arn
Huttmann, GPRG, University of Sheffield. Units are metres for depth, mg l1 for concentrations.
Since the molecular weights of ammonium and nitrate are 18 and 62 (g mole1 ), respectively, then
1 = 1 mmol NH+ l1 , 1 mg NO l1 = 1 mmol NO l1 . Nitrate shows a sharp
1 mg NH+
4 l
4
3
3
18
62
spike (of about 10 mmol l1 ) at a depth of 19 m. There appears to be a second front at 23 m: the
nitrate produced at 19 m diffuses there and takes out (according to (7.201)) either the inorganic
carbon CH2 O or the acid H+

where the parameters i are defined analogously to (7.194), i.e.,


i =

ci
cNH+

(7.205)

ci being the scale for ci . From the values given above, we can suppose that all the
i  1.
We now mirror the Four Ashes discussion, supposing that 1. The reaction
front at the plume fringe should then be thin as before, with oxygen diffusing to
the front from outside the plume, and ammonium diffusing to it from inside, and
both concentrations being zero at the front. The fringe position is unknown, and we
seek a conservation law to provide an extra condition for it analogous to (7.197).
Denote the right hand sides of (7.202) by e1 e5 . There are five reactants but only
three reactions, therefore there are two (linear) relationships between
 the ei : these
provide suitable jump conditions across the fringe. The equations i i ei = 0 are
satisfied for any 1 and 2 provided 3 = 1 + 12 2 , 4 = 1 + 34 2 , and 5 = 2 .
Selecting (1 , 2 ) = (1, 0) and (0, 1), we thus have
e1 e3 + e4 = 0,
1
3
e2 + e3 + e4 e5 = 0,
2
4

(7.206)

and it follows from this that in consideration of (7.204) there are two conserved
quantities across the fringe (i.e., flux in equals flux out); explicitly (with the same

7.8 Three Specific Remediation Problems

431

caveat concerning the normal n as expressed in (7.199)), we have


 c

NH+
4

NO
3

cNO
3

c +
+ H+ H
n

= 0,



cNO
1
3
cO2
c +
cCH2 O +
3
+ NO
+ H+ H CH2 O
O2
= 0,
3
n
2
n
4
n
n

(7.207)

where [j ]+
denotes the jump in j across the fringe.
Consulting (7.201), we can write the dimensionless reaction rates ri in terms of
Monod rates
ci
(7.208)
Mi =
Ki + ci
in the following way:
r1 = MNH+ MO2 ,
4

r2 = k2 MNO MH+ MCH2 O ,


3

r3 = k3 MNH+ MNO ,
4

(7.209)
where k2 and k3 are dimensionless constants (the ratios of the saturated maxima
kmax of reactions 2 and 3 in (7.201) to that of reaction 1). The requirement that reaction rates vanish on either side of the fringe is satisfied if cNH+ = 0 outside the
4
plume (then r1 = r3 = 0), cO2 = 0 within the plume (so r1 = 0) and if cNO = 0 on
3
both sides (then r2 = r3 = 0). This simplifies the situation to one where pre-existing
soil nitrate concentrations are low, and assumes all the nitrate produced in the reaction front by reaction 1 is consumed there by reaction 2. An alternative would be
that the produced nitrate diffuse away on either side, but this is not consistent with
a fast reaction 2 unless there is a natural source for nitrate production (e.g., from
cNH+

surface composting).11 Since from (7.203), H+ 105 , this suggests that n 4


is approximately continuous across the fringe, which is thus diffusive in character
(providing Pe 1). In that case the fringe is located simply by advection of ammonium.
The second jump condition then provides a flux boundary condition for cCH2 O ,
which is not constrained to be zero at the front. In addition, existence of a spike of nitrate within the front must be determined by solution of the reaction equations within
the front. The reaction front is of thickness (1/Pe)1/2 (taking Pe = Pe for simplicity), and within this front, the nitrate concentration is given approximately by
2 cNO
3

N 2
11 Alternatively,

+ NO (r1 r2 0.6r3 ) = 0,
3

there may be more than one reaction front. This is suggested by Fig. 7.9.

(7.210)

432

7 Groundwater Flow

where N is a suitably rescaled normal variable. We can expect this to be solvable for cNO subject to cNO 0 as N , since for example if we lin3
3
earise the nitrate Monod coefficient, MNO = cNO , then the solution can be written
3
3
using a Greens function which depends also on all the other reactant concentrations.
There remains the issue of determining a boundary condition for H+ at the front.
Counting conditions, the first flux condition in (7.207) determines the location of the
fringe. We already have boundary conditions for cNH+ and cO2 (both equal zero),
4
and the nitrate spike is determined from (7.210). Thus, in principle we know the
jump in flux of ammonium, oxygen and nitrate, and by integrating (7.210) and its
equivalents for ammonium and oxygen through the front, we know that


(r1 + r3 ) dN =


O2


NO
3

 c


2r1 dN =

NO
3

+
,

N
cO2
N

+
,

(7.211)

(r1 r2 0.6r3 ) dN = 0.


Thus we know the values of ri dN for i = 1, 2, 3 and this tells us (by integrating

through the reaction front) the values of ri dN for i = 4, 5, and thus the jump in
c

H+
flux of N
; this provides the extra boundary condition we seek. The jump in flux

of cCH2 O is also given by 1.25CH2 O r2 dN , but this is equivalent to (7.207)2 . In
this way, the approximate model provides all the conditions necessary to determine
the solution.

St. Albans At St. Albans, there has been a petroleum spillage (at a filling station)
into an underlying chalk aquifer. The fluids are LNAPLs: hydrocarbons, BTEX (a
cancer-forming aromatic hydrocarbon12 and MTBE.13 BTEX is retarded compared
to MTBE and thus forms a secondary plume within the MTBE plume. The LNAPLs
have seeped through the unsaturated zone and sit on top of the chalk aquifer, acting as a source (via dissolution) of contaminant to the underlying groundwater
flow.
The sequence of reactions appears to be similar to those of the other examples,
with oxidation by oxygen and nitrate at the plume fringe, and by Mn (manganese),
Fe (iron) and SO2
4 (sulphate) in the plume core. The sequence of reactions which
12 More specifically, BTEX refers to a suite of volatile hydrocarbons, the acronym referring to benzene, toluene, ethylbenzene and xylene, with chemical formulae C6 H6 (benzene), C7 H8 (toluene),
C8 H10 (ethylbenzene and xylene); we use toluene in the chemical reaction model.
13 Methy

tert-butyl ether, C5 H12 O.

7.8 Three Specific Remediation Problems

433

are modelled is the following:


r1

C7 H8 + 9O2 + 3H2 O 7CO2


3 ,
r2

+
C7 H8 + 7.5H2 O 2.5CO2
3 + 4.5CH4 + 5H ,
r3

2+
+ 50H+ ,
C7 H8 + 36Fe3+ 7CO2
3 + 36Fe
r4

+
CH4 + 2O2 CO2
3 + 2H + H2 O,

CH4 + 8Fe

3+

r5

+ 3H2 O

CO2
3

+ 8Fe

2+

(7.212)
+

+ 10H ,

r6

Fe2+ + O2 + H+ Fe3+ + 0.5H2 O,


r7

C7 H8 + 36FeOOH(s) + 58H+ 36Fe2+ + 7CO2


3 + 51H2 O,
r8

CH4 + 8FeOOH(s) + 7H+ 8Fe2+ + CO2


3 + 5H2 O.
Maximum reaction rates vary over about four orders of magnitude, from about
1012 mol l1 s1 (reactions 2, 3, 7) to 108 mol l1 s1 (reaction 8). The principal oxidising reaction rate r1 1010 mol l1 s1 . Contaminant levels are of order
107 mol l1 , with oxygen level of order 106 mol l1 . Plume depth and length are
d 20 m, l 300 m, dispersivity parameters are taken as  1 m, 0.2 m.
The hydraulic conductivity of the upper chalk aquifer is 25 m d1 (that of the lower
part is only about 0.1 m d1 ), and the hydraulic gradient is 1.75 103 , thus the
longitudinal flux scale is U 16 m y1 . With these values, the parameters defined
by (7.186) have values
Pe 300,

Pe 10,

106 .

(7.213)

As now seems monotonously to be the case, reactions are fast, longitudinal dispersion is small, but transverse dispersion may be effective because of the high aspect
ratio l/d.
The distinguishing feature of this particular site is that, although the chalk is very
porous (  0.3), it has a very low effective permeability, presumably due to chemical adsorption by the chalk. On the other hand, the chalk matrix is dissected into
blocks by numerous fractures, and thus acts as a dual porosity system. Contaminant
can diffuse into the pore space of the matrix, and the issue of concern is whether and
how fast this happens, since storage in matrix blocks will act as a residual source of
contamination after the fracture system has been flushed.
We have considered this problem before, in Sect. 7.6.1. Let us suppose for simplicity that dispersivities of matrix and fractures are constant and isotropic but not
necessarily equal. (Additionally, we suppose the local fracture dispersivity Df is
equal to the macroscopic fracture dispersivity DF .) The point forms of the equations
describing a single reactant are given by (7.165), and the block scale and averaged

434

7 Groundwater Flow

equations for the matrix concentration are given by (7.167) and (7.170):


cm
1
+ u m .x cm = x2 cm + Pem Sm + 2 n. X cm |M ,
Pem
t

(7.214)


cm
2
2
Pem
+ um .X cm = X cm + Pem Sm ,
T
where
Pem =

Ul
,
Dm

S0 l
.
U c0

(7.215)

Analogous dimensionless forms for the fracture average and block scale equations
are, from (7.162) and (7.165)2 ,


cf
Dm
+ u f .x cf = x2 cf + Pef Sf 2
Pef
n.X cm |M ,
t
Df f
(7.216)


cf
D
m
(n. X cm )|M ,
+ uf . X cf = X2 cf + 2 Pef Sf
Pef
T
Df f
and Pef = U l/Df .
The question is how the interfacial transport term should be modelled. In principle we solve the block equation for cm with cm = cf on M, yielding the interfacial
term as an integral convolution of cf . Putting this into the block equation for cf and
solving this then determines cf and thus gives the interfacial term, which closes the
description of the averaged equations.
The upshot of our earlier discussion was that the details of the homogenisation
process depend on the relation between the parameters = dB / l, Pe and . Classical homogenisation theory as in Sect. 7.6.1 assumes all the parameters are O(1)
apart from , but this is unlikely ever to be appropriate. Estimates for St. Albans
may be Pef 102 , 106 , 104 ; in addition, um  uf and Dm  Df . Thus
the reaction terms in the macroscopic equations are always large, and this suggests
that, as before, reactions will be restricted to thin fronts. A complication in (7.214)
is that the interfacial transport term may be large also, so this now needs to be determined.
To be specific, let us consider a simple two component reaction similar to (7.191),
i.e.,
r

C7 H8 + O2 products,

(7.217)

and let p denote dimensionless BTEX concentration and c denote dimensionless


oxygen concentration. Just as in (7.192), we can put Sm = S in (7.221), and there
is a corresponding reaction term S in the equivalent equation for p: S can be taken
as the product of Monod rates in (7.193), and for simplicity we take the linear rates,
thus S = pc.

7.8 Three Specific Remediation Problems

435

The critical parameter in (7.214)2 is 2 Pem ; we write


Bm = 2 Pem ,

Bf = 2 Pef ;

(7.218)

if Pef 102 then Bf O(1), and Bm 1: reaction in the block is fast.


Just as in (7.196), cm pm satisfies an advective-diffusion conservation equation in the block, diffusion acting on a time scale t 2 Pem , where t is the macroscopic time variable. The parameter 2 Pem is crucial. If we suppose that molecular diffusion applies in the blocks, then Dm 109 m2 s1 , Pem 105 , and
2 Pem 103 is small. Therefore on the long macroscopic time scale, diffusion
smoothes cm pm , and we can take it to be locally constant in the blocks (and thus
also the fractures). Since reaction is fast in the blocks, this implies
cm pm ,

(7.219)

and therefore the block reaction rate for cm can be written as


2
.
2 Pem Sm Bm cm

(7.220)

When Bm is large there is a reaction boundary layer like a rind at the block
surface, within which cm satisfies
2 cm
2
Bm c m
0.
N 2

(7.221)

In the interior of a block inside the plume the oxygen is depleted. and we have
cm 0 as N . The first integral of (7.221) determines the flux to the blocks
as

cm
2Bm 3/2
=
c ,
(7.222)
N
3 f
using the fact that cm = cf on M. The matrix concentration is negligible, and the
average fracture concentration satisfies the Eq. (7.216)1

Pef



cf
B
2Bf P ef 3/2
1
f
c ,
+ u f . x cf = x2 cf 2 cf2 2
t

f
3Pem f

(7.223)

where we take Sf = cf2 as in the matrix, and we have supposed that


cf2 = cf2 ,

3/2

3/2

cf = cf .

(7.224)
B

With Bf O(1), Pef 102 , Pem 105 , the flux term to the blocks, 2f cf2 1 and

2Bf P ef 3/2
the fracture reaction term 21
3Pem cf 1; the implication is that the fracture
f
concentration of oxygen within the plume rapidly decreases within both blocks and

436

7 Groundwater Flow

fractures to very small levels. The dual porosity appears to have little effect on the
characteristics of the reactant distributions.

7.9 Precipitation and Dissolution


A sedimentary basin, as illustrated in Fig. 7.10, refers to an accumulation of sediments (of typical depth 10 km) derived from river outwash sands and silts, or marine
deposited muds and microfossils. Sedimentary basins are everywhere in continental rocks: we have the Paris basin, the London basin, the North Sea, and so on.
As sediments accumulate and are buried, they are subjected to increasing heat and
pressure, and these two factors enable the formation of intergranular cements, which
thus convert the sediments to rock (sand to sandstone; clay to shale, marine organisms to limestone). As they are buried, the sediments also compact, expelling pore
water. Depending on the permeability, this can lead to pore pressures above hydrostatic, a situation which is of concern in oil-drilling operations (and is discussed
further in Sect. 7.11).
Diagenesis refers generally to the process of chemical alteration to rock, and in
this section we study the effects of diagenesis on the groundwater flow of accumulating sediments within a sedimentary basin. The particular type of diagenesis which
we discuss is the conversion of smectite (a form of hydrated clay) to illite (a dehydrated clay) via a dewatering reaction. The resultant release of water is also a potential cause of excess pore pressures, but the main purpose of the present discussion
is to show how the use of an approximation which we may call the weak solubility
limit allows enormous simplification of quite complicated reaction schemes.
One view of the smectiteillite reaction is to treat it using first order kinetics, thus
SS IS + nH2 O,

(7.225)

where each mole conversion yields n moles of water: S denotes smectite, I denotes
illite, and the superscript S denotes the solid phase (likewise, L will denote an aqueous phase). Such a scheme is not inconsistent with at least some experimental data,
and the rate factor involved depends on temperature, with an activation energy in
the range 6080 kJ mol1 .
However, it is likely that the transformation of smectite to illite occurs through
a compound sequence of precipitation and dissolution reactions; one possible deFig. 7.10 Schematic of a
sedimentary basin. Sediment
accumulates from outflow
from rivers, and also through
the settlement of marine
organisms

7.9 Precipitation and Dissolution

437

scription is the following:


R1

SS XL + nH2 O,
R2

L
KFs K+L + AlOL
2 + s SiO2 ,
R3

L
S
L
K+L + AlOL
2 + f X f I + SiO2 ,
R4+
L 
SiO2
R4

(7.226)

Qz.

The smectite SS dissolves to form a hydrous silica combination XL , such as


Si4 O10 (OH)2 . Additionally, illite precipitation requires potassium ions, and these
may be obtained from the dissolution of potassium feldspar (in the second reaction);
the aluminium hydroxyl ions AlOL
2 act in the same way. The hydrous silica combination now combines with the potassium and aluminium to form illite precipitate,
together with aqueous silica SiOL
2 , which itself precipitates as quartz Qz. Taking
suitable multiples of the reactions and adding to eliminate the aqueous phases, the
overall reaction is found to be
S + f 1 KFs I + nH2 O + f 1 (s + 1)Qz.
R

(7.227)

Ideally we would like to be able to write kinetics for (7.227) analogously to (7.225),
with a recipe for the effective reaction rate R. Note that all the reactions in (7.226)
are precipitation or dissolution reactions, and therefore the reaction rates are proportional to grain surface area.
It turns out, at least for this reaction scheme (but we might suspect more generally), that the weak solubility limit allows such a recipe to be found. To illustrate the
method, note that conservation equations for the concentrations S, X, F , K, A, L,
Q, I of the substances smectite, hydrous silica, feldspar, potassium ions, aluminium
hydroxyl ions, aqueous silica, quartz and illite satisfy equations of the type

(1 )S + . (1 )Sus = r1 ,
t

(X) + . Xul .(D.X) = r1 f r3 ,


t

(7.228)

and so on, where the reaction rates ri are scaled with respect to the surface rates
Ri by the specific interfacial surface areas i (thus ri = i Ri ). ul and us are the
liquid and solid Darcy fluxes (i.e., the volume fluxes per unit area). We allow a
non-zero solid velocity in order to cater for the effects of compaction of the porous
matrix. There are six other equations of the type in (7.228), together with a water
conservation equation (this is the equation for ). Of the total of nine equations,
four are for aqueous concentrations. That for A is identical to that for K, and we
ignore it henceforth.

438

7 Groundwater Flow

The weak solubility limit is associated with the observations that the aqueous
dissolved species X (hydrous silica), K (potassium) and L (silica) are typically
present in trace quantities of the order of 10100 ppm (1 ppm = 103 kg m3 ),14
and thus the concentrations of the aqueous phases are much less than those of the
solid phases. When the model is suitably non-dimensionalised, the result is that the
transport terms for the aqueous phases are very small, so that the corresponding
reaction terms can be taken to be in equilibrium. The reaction terms in the equations
for X, K and L are, respectively, r1 f r3 (as above), r2 r3 and sr2 + r3 r4+ + r4 .
From this, we obtain the three relationships
r1 f r3 ,
r2 r3 ,

(7.229)

r4+ r4 (s + 1)r3 .
Since the rate of dissolution of S is r1 and the rate of precipitation of I is f r3 ,
this immediately shows that first order kinetics of the form (7.225) does apply, with
the reaction rate being f r3 . Since r4+ is a precipitation rate, and r4 the dissolution
+

rate of the same mineral pair SiOL


2 Qz, and since either r4 = 0 or r4 = 0, it is
+

apparent that (7.229)3 determines r4 and r4 together. Thus all the reaction rates
can be written in terms of r3 , and this then determines the aqueous phase pseudoequilibrium concentrations of X, L and K in terms of r3 and various temperature
dependent rate factors, since the kinetic rates ri are prescribed in terms of these.
If the smectite equation (7.228)1 is written in terms of smectite volume fraction
S , then it becomes

S
MS
f r3 ,
+ . S us =
t
S

(7.230)

where MS is the molecular weight of smectite, and S its density. An equivalent


equation for illite is

MI
I
r3 ,
+ . I us = f
t
I

(7.231)

and there are two similar equations for F and Q (with the right hand sides being
proportional to r3 ). In addition, the porosity satisfies

Mw
+ . ul = n
f r3 .
t
w

(7.232)

It remains to determine the rate constant r3 in terms of the reactant concentrations.


1 ppm = 1 mg kg1 , but 1 m3 of water weighs 103 kg, so for aqueous solutions, this is
equivalent to 103 kg m3 .
14 Actually,

7.9 Precipitation and Dissolution

439

We assume that the reaction rates take the general form (in which D denotes
dissolution, and P precipitation)


ci
,
riD = i Ri 1
cis +
(7.233)


ci
P
r i = i R i
1 ,
cis
+
where ci is the relevant aqueous concentration of phase i and cis is the associated
solubility limit, i.e., the saturation concentration of aqueous phase i in the presence
of solid phase s. The rate factor Ri generally depends on temperature. (7.233) states
that precipitation occurs when a solution is oversaturated, and dissolution occurs if
it is undersaturated.
For the specific case of the precipitation/dissolution scheme in (7.226), we suppose that
r1D = S RSD [1 X ]+ ,
r2D = f RFD [1 K ]+ [L L ]+ ,
r3P = I RIP [K K ]+ [X X ]+ ,

(7.234)

r4P = Q RQP [L 1]+ ,


r4D = Q RQD [1 L ]+ .
In these equations we have assumed that specific surface area is proportional to
volume fraction, thus S = S /dp , and we have absorbed the grain size dp into
the reaction rates. The reaction rate subscripts describe what they represent: SD,
smectite dissolution; FD, feldspar dissolution; IP, illite precipitation; QP, quartz
precipitation; QD, quartz dissolution.
The quantities D represent dimensionless aqueous concentrations (of phase D),
scaled with a suitable solubility limit: cX is scaled with cXS , the solubility limit for
hydrous silica in the presence of smectite; cL is scaled with cLQ , the solubility
limit for silica in the presence of quartz. Potassium is more complicated, because
in (7.226), we see that dissolution of feldspar produces two aqueous phases, and
precipitation of illite requires two. We can in fact define four further solubility limits:
cLF , that of silica in the presence of feldspar; cKF , that of potassium in the presence
of feldspar; cKI , that of potassium in the presence of illite; and cXI , that of hydrous
silica in the presence of illite. The assumption is then that, for example, feldspar will
dissolve only if both potassium and silica are undersaturated, i.e., cK < cKF and
cL < cLF . In writing (7.234), we have scaled cK with cKF , and the three solubility
ratios in (7.234) are therefore defined by
L =

cLF
,
cLQ

K =

cKI
,
cKF

X =

cXI
.
cXS

(7.235)

440

7 Groundwater Flow

Various possibilities can now occur depending on the values of the different solubility limits. We suppose that the solubility of hydrous silica X with respect to illite
is much less than that with respect to smectite (thus X  1), and we suppose this
is also true for potassium, that is to say, K  1. On the other hand, we suppose
that the solubilities of silica with respect to feldspar or quartz are comparable, so
that L 1, and in fact we will take L > 1 as appears to be appropriate (indeed
otherwise the smectiteillite transition will not occur in this model).
The equilibrium equations (7.229) are
S RSD [1 X ]+ = f r3 ,
F RFD [1 K ]+ [L L ]+ = r3 ,
Q RQP [L 1]+ Q RQD [1 L ]+ = (s + 1)r3 ,

(7.236)

r3 = I RIP [K K ]+ [X X ]+ ,
and these are four equations for the reaction rate r3 and the three aqueous concentrations X , L and K . Assuming non-zero reaction rates, and taking K , X  1,
and L > 1, we deduce from (7.236) that L > 1 (thus quartz precipitates from
solution), and
r3 = I RIP K X ,
X = 1

f r3
,
S RSD

L = 1 +

(s + 1)r3
,
Q RQP

K = 1

r3

F RFD L 1

(7.237)
(s+1)r3 
Q RQP

whence the basic reaction rate r3 is determined by




f r3
r3
r3 = I RIP 1
1

S RSD
F RFD L 1


.

(s+1)r3

(7.238)

Q RQP

The right hand side of this expression is a decreasing function of r3 if it is positive,


while the left hand side is increasing; therefore this expression defines the rate r3
uniquely in terms of the solid phase concentrations of feldspar, quartz, smectite and
illite.
A general complication in solving for S and the other reactant porosities is that
r3 depends on S , I , Q and F . Given us , (7.230) is a hyperbolic equation for S
with boundary condition S = S0 on the upper surface z = h of a sedimentary basin
b < z < h. The equations for the solid fractions Y , Y = S, I, Q, F are all of the
form


Y
+ . Y us = Y r3 ,
(7.239)
t

7.10

Consolidation

441

for certain constants Y , and if Y = Y0 on z = h, then Y can be written as a linear


combination of S and I ,
Y =

(Y S0 + Y0 )I + (I Y0 Y I0 )S
I S0 + I0

(7.240)

Therefore, the reaction rate r3 can generally be written explicitly as a function of S


and I , and the diagenesis model collapses to equations for S , I and , together
with Darcys law.
Diagenesis (at least in this theory) turns out to have a minor quantitative effect
on groundwater flow, essentially because the source term in (7.232) is relatively
small. What is perhaps of more interest is that a fairly complicated sequence of
precipitation/dissolution steps can be reduced, in the limit of weak solubility, to a
model with first order kinetics, albeit with a complicated (but explicitly defined)
reaction rate. In fact, this observation is likely to be true in general. Suppose we
have a sequence of precipitation and dissolution steps for solids Si and liquids Lj :
R1

L1 + S1 + ,
R2

(7.241)

S2 + L2 + .
Each reaction step necessarily involves at least one aqueous phase component, and
thus all the reaction rates R1 , . . . , Rn occur in the conservation equations for the
aqueous phase components. Since these can all be taken to be in equilibrium, then
if there are k different aqueous phase components, we obtain k relations for the n
reactions. If k = n 1, then all the reaction rates can be written in terms of the
overall production rate, and first order kinetics will apply.
In the present example (7.226), there are five reaction steps, and three aqueous
components (lumping K+L and Al(OH)L
4 together), but the precipitation/dissolution of quartz is effectively one reaction (either but not both at once can occur),
and so the condition n = k + 1 is effectively met. More generally, we see that the
production of solid precipitate P from solid substrates S through a sequence of
intermediate dissolution/precipitation steps may often lead to this situation.

7.10 Consolidation
Consolidation refers to the ability of a granular porous medium such as a soil to
compact under its own weight, or by the imposition of an overburden pressure. The
grains of the medium rearrange themselves under the pressure, thus reducing the
porosity and in the process pore fluid is expelled. Since the porosity is no longer
constant, we have to postulate a relation between the porosity and the pore pressure p. In practice, it is found that soils, when compressed, obey a (non-reversible)
relation between and the effective pressure
pe = P p,
where P is the overburden pressure.

(7.242)

442

7 Groundwater Flow

Fig. 7.11 Form of the


relationship between porosity
and effective pressure. A
hysteretic decompression-reconsolidation loop is
indicated. In soil mechanics
this relationship is often
written in terms of the void
ratio e = /(1 ), and
specifically
e = e0 Cc log pe , where Cc
is the compression index

The concept of effective pressure, or more generally effective stress, is an extremely important one. The idea is that the total imposed pressure (e.g., the overburden pressure due to the weight of the rock or soil) is borne by both the pore fluid and
the porous medium. The pore fluid is typically at a lower pressure than the overburden, and the extra stress (the effective stress) is that which is applied through grain
to grain contacts. Thus the effective pressure is that which is transmitted through
the porous medium, and it is in consequence of this that the medium responds to the
effective stress; in particular, the characteristic relation between and pe represents
the nonlinear pseudo-elastic effect of compression.
As pe increases, so decreases, thus we can write (ignoring irreversibility)
pe = pe (),

pe () < 0.

(7.243)

Taking the fluid density to be constant, we obtain from the conservation of mass
equation the nonlinear diffusion equation



k()  
(7.244)
pe () ,
t = .

assuming Darcys law with a permeability k, ignoring gravity, and taking P as constant. This is essentially the same as the Richards equation for unsaturated soils.
The dependence of the effective pressure on porosity is non-trivial and involves
hysteresis, as indicated in Fig. 7.11. Specifically, a soil follows the normal consolidation line providing consolidation is occurring, i.e. pe > 0. However, if at some
point the effective pressure is reduced, only a partial recovery of takes place.
When pe is increased again, more or less retraces its (overconsolidated) path to
the normal consolidation line, and then resumes its normal consolidation path. Here
we will ignore effects of hysteresis, as in (7.243).
When modelling groundwater flow in a consolidating medium, we must take
account also of deformation of the medium itself. In turn, this requires prescription
of a constitutive rheology for the deformable matrix. This is often a complex matter,
but luckily in one dimension, the issue does not arise, and a one-dimensional model
is often what is of practical interest. We take z to point vertically upwards, and let v l

7.10

Consolidation

443

and v s be the linear (or phase-averaged) velocities of liquid and solid, respectively.
Then ul = v l and us = (1 )v s are the respective fluxes, and conservation of
mass of each phase requires

Darcys law is then

(v l )
+
= 0,
t
z
{(1 )v s }
+
= 0;

t
z

(7.245)





k p
vl v s =
+ l g ,
z

(7.246)

while the overburden pressure is


P = P0 + s (1 ) + l g(h z);

(7.247)

here z = h represents the ground surface and P0 is the applied load. (7.247) assumes
variations of are small. More generally, we would have P /z = [s (1 ) +
l ]g. The effective pressure is then just pe = P p.
We suppose these equations apply in a vertical column 0 < z < h, for which
suitable boundary conditions are
vl = v s = 0
p = 0,

at z = 0,
h = v s

at z = h,

(7.248)

and with an initial condition for p (or ).


The two mass conservation equations imply
vs =

v l
.
1

Substituting this into (7.246), we derive, using (7.245),





p

k
=
(1 )
+ l g .
t
z
z

(7.249)

(7.250)

If we assume the normal consolidation line takes the commonly assumed form (see
Fig. 7.11)

= e0 Cc ln pe /pe0 ,
(7.251)
1
then we derive the consolidation equation



pe
pe
k
pe
=
(1 )
+ (1 )g ,
t
z
Cc (1 )2 z
where  = s l .

(7.252)

444

7 Groundwater Flow

If Cc is small (and typical values are in the range Cc 0.1) then varies little,
and the consolidation equation takes the simpler form
pe
2 pe
= cv 2 ,
t
z

(7.253)

where
cv =

k
pe
Cc (1 )

(7.254)

is the coefficient of consolidation.


Suitable boundary conditions are
pe
+ (1 )g = 0 at z = 0,
z
pe = P0

(7.255)

at z = h,

and if the load is applied at t = 0, the initial condition is


pe = (1 )g(h z)

at t = 0.

(7.256)

The equation is trivially solved. The consolidation time is


tc

h2 Cc (1 )h2
=
,
cv
kpe

(7.257)

and depends primarily on the permeability k. If we take k 1014 m2 (silt),


Cc = 0.1, = 0.3, = 103 Pa s, P0 = 105 Pa (a small house), then cv
105 m2 s1 , and te 1 year for h 10 m.

7.11 Compaction
Compaction is the same process as consolidation, but on a larger scale. Other mechanisms can cause compaction apart from the rearrangement of sediments: pressure
solution in sedimentary basins, grain creep in partially molten mantle (see Chap. 9).
The compaction of sedimentary basins is a problem which has practical consequences in oil-drilling operations, since the occurrence of abnormal pore pressures
can lead to blow-out and collapse of the borehole wall. Such abnormal pore pressures (i.e., above hydrostatic) can occur for a variety of reasons, and part of the
purpose of modelling the system is to determine which of these are likely to be
realistic causes. A further distinction from smaller scale consolidation is that the
variation in porosity (and, particularly, permeability) is large.
The situation we consider was shown in Fig. 7.10. Sediments, both organic and
inorganic, are deposited at the ocean bottom and accumulate. As they do so, they
compact under their weight, thus expelling pore water. If the compaction is fast (i.e.,

7.11

Compaction

445

the rate of sedimentation is greater than the hydraulic conductivity of the sediments)
then excess pore pressure will occur.
Sedimentary basins, such as the North Sea or the Gulf of Mexico, are typically
hundreds of kilometres in extent and several kilometres deep. It is thus appropriate
to model the compacting system as one-dimensional. A typical sedimentation rate
is 1011 m s1 , or 300 m My1 , so that a 10 kilometre deep basin may accumulate
in 30 My (30 million years). On such long time scales, tectonic processes are important, and in general accumulation is not a monotonic process. If tectonic uplift
occurs so that the surface of the basin rises above sea level, then erosion leads to denudation and a negative sedimentation rate. Indeed, one purpose of studying basin
porosity and pore pressure profiles is to try and infer what the previous subsidence
history wasan inverse problem.
The basic mathematical model is that of slow two-phase flow, where the phases
are solid and liquid, and is the same as that of consolidation theory. The effective pressure pe is related, in an elastic medium, to the porosity by a function
pe = pe (). In a soil, or for sediments near the surface up to depths of perhaps
500 m, the relation is elastic and hysteretic. At greater depths, more than a kilometre, pressure solution becomes important, and an effective viscous relationship
becomes appropriate, as described below. At greater depths still, cementation occurs and a stiffer elastic rheology should apply.15 In addition, the permeability is a
function k = k() of porosity, with k decreasing to zero fairly rapidly as decreases
to zero.
Let us suppose the basin overlies an impermeable basement at z = 0, and that its
surface is at z = h; then suitable boundary conditions are
v s = v l = 0 at z = 0,
h = m
s + vs
pe = 0,

at z = h,

(7.258)

s is the prescribed
where v s and v l are solid and liquid average velocities, and m
sedimentation rate, which we take for simplicity to be constant.
If we assume a specific elastic compactive rheology of the form


(7.259)
pe = p0 ln(0 /) (0 ) ,
p0
then non-dimensionalisation (using a depth scale d = (s
and a time scale mds )
l )g
and simplification of the model leads to the nonlinear diffusion equation, analogous
to (7.250),




1

k(1 )2
=
1 ,
(7.260)
t
z
z

where the permeability is defined to be

k = k0 k(),
k0 being a suitable scale for k.
15 Except

that at elevated temperatures, creep deformation will start to occur.

(7.261)

446

7 Groundwater Flow

The dimensionless parameter is given by


=

K0
,
m
s

(7.262)

where K0 = k0 (s l )g/ is essentially the surface hydraulic conductivity, and


we can then distinguish between slow compaction (  1) and fast compaction
( 1). Typical values of depend primarily on the sediment type. For m
s =
1011 m s1 , we have 0.1 for the finest clay, 109 for coarse sands. In general, therefore, we can expect large values of . The associated boundary conditions
for the model become
z = 0 at z = 0,
= 0 ,



) 1 1
h = 1 + k(1
at z = h.
z

(7.263)

Slow Compaction,  1 When is small, overpressuring occurs. A boundary


layer analysis is easy to do, and shows that 0 in the bulk
of the (uncompacted)
sediment, while a compacting boundary layer of thickness t exists at the base.
Fast Compaction,  1 The more realistic case of fast compaction is also the
more mathematically interesting. Most simply, the solution when 1 is the equilibrium profile
= 0 exp[h z];

(7.264)

the exponential decline of porosity with depth is sometimes called an Athy profile,
but it only applies while k 1. If we assume a power law for the dimensionless
permeability of the form
k = (/0 )m ,
then we find that k reaches one when decreases to a value


1
= 0 exp ln ,
m

(7.265)

(7.266)

and this occurs at a dimensionless depth


1
ln
m

(7.267)

0 (1 e )
.
1 0

(7.268)

=
and time
t =

Typical values m = 8, = 100, 0 = 0.5, give values = 0.28, = 0.58, t =


0.71. In particular, for a reasonable depth scale of 1 km (corresponding to p0 =
2 107 Pa = 200 bars), this would correspond to a depth of 580 m. Below this, the

7.11

Compaction

447

Fig. 7.12 Solution of (7.260)


for = 100 at times
t = t 0.71 and at t = 2.
The porosity (horizontal axis)
is plotted as a function of the
scaled vertical height z/ h(t).
The solid lines are numerical
solutions, whereas the dotted
lines are the large
equilibrium profiles. There is
a clear divergence at depth for
t > t

profile is not equilibrated, and the pore pressure is elevated. Figure 7.12 shows the
resulting difference in the porosity profiles at t = t and t > t , and Fig. 7.13 shows
the effect on the pore pressure, whose gradient changes abruptly from hydrostatic to
lithostatic at the critical depth.

Fig. 7.13 Hydrostatic,


overburden (lithostatic) and
pore pressures at t = 5 and
= 100, as functions of the
scaled height z/ h(t). The
transition from equilibrium to
non-equilibrium compaction
at the critical depth is
associated with a transition
from normal to abnormal
pore pressures. The dashed
lines represent two distinct
approximations to the pore
pressure profile, respectively,
valid above and below the
transition region

448

7 Groundwater Flow

If we take = O(1) and 1, then formally m 1, and it is possible to


analyse the profile below the critical depth. One finds that



1

= exp ln m + O(1) ,
(7.269)
m
which can explain the flattening of the porosity profile evident in Fig. 7.12, and
which is also seen in field data.
Viscous Compaction Below a depth of perhaps a kilometre, pressure solution
at intergranular contacts becomes important, and the resulting dissolution and local reprecipitation leads to an effective creep of the grains (and hence of the bulk
medium) in a manner analogous to regelation in ice. For such viscous compaction,
the constitutive relation for the effective pressure becomes
pe = .us .

(7.270)

In one dimension, the resulting dimensionless model is



+
(1 )u = 0,
t
z


p
u = k
+1 ,
z

p =

(7.271)

u
,
z

where p is the scaled effective pressure. The compaction parameter is the same as
before, and the extra parameter can be taken to be of O(1) for typical basin depths
of kilometres. Boundary conditions for (7.271) are
u=0
p = 0,

on z = 0,
= 0 ,

h = 1 + u at z = h.

(7.272)

This system can also be studied asymptotically. When  1, compaction is slow


and a basal compaction layer again forms. When 1, explicit solutions can again
be obtained. There is an upper layer at equilibrium, but now the porosity decreases
concavely with depth.16 As before, there is a transition when = , and below this



2
= exp ln m + O(1) ,
(7.273)
m
similar to (7.269).
16 In

view of Chap. 6, we need to be careful here. The function is mathematically concave, i.e., the
rate of decrease of porosity with depth increases as depth increases.

7.12

Notes and References

449

Fig. 7.14 Evolution of the


porosity as a function of
depth h z, with a viscous
rheology, at = 100. The
upper concave part is in
equilibrium, while
overpressuring occurs where
the profile is flatter below this

The main distinction between viscous and elastic compaction is thus in the form
of the rapidly compacted equilibrium profile near the surface (Fig. 7.14). The concave profile is not consistent with observations, but we need not expect it to be, as
the viscous behaviour of pressure solution only becomes appropriate at reasonable
depths. A more general relation which allows for this is a viscoelastic compaction
law of the form
.us =

1 dpe pe
.
Ke dts

(7.274)

7.12 Notes and References


Flow in porous media is described in the books by Bear (1972) and Dullien (1979).
More recent versions, for example by Bear and Bachmat (1990) have developed a
taste for more theoretical, deductive treatments based on homogenisation (see below) or averaging, with a concomitant loss of readability. The classic geologists
book on groundwater is by Freeze and Cherry (1979) and the classic engineering
text is by Polubarinova-Kochina (1962). A short introduction, of geographical style,
is by Price (1985). A more mathematical survey, with a variety of applications, is by
Bear and Verruijt (1987). The book edited by Cushman (1990) contains a wealth of
articles on topics of varied and current interest, including dispersion, homogenisation, averaging, dual porosity models, multigrid methods and heterogeneous porous
media. Further information on the concepts of soil mechanics can be found in Lambe
and Whitman (1979).

450

7 Groundwater Flow

Homogenisation The technique of homogenisation is no more than the technique of averaging in the spatial domain, most often formulated as a multiple scales
method. Whole books have been written about it, for example those by Bensoussan
et al. (1978) and Sanchez-Palencia (1983). For application to porous media, see, for
example, Enes article in the book edited by Cushman (1990).
Piping Many dams are built of concrete, and in this case the problems associated
with seepage do not arise, owing to the virtual impermeability of concrete. Earth
and rockfill dams do exist, however, and are liable to failure by a mechanism called
piping. The Darcy flow through the porous dam causes channels to form by eroding
away fine particles. The resultant channelisation concentrates the flow, increasing
the force exerted by the flow on the medium and thus increasing the erosion/collapse
rate of the channel wall. We can write Darcys law as a force balance on the liquid
phase,

(7.275)
vl l gk
k
(k being vertically upwards) and vl /k is an interactive drag term; then the corresponding force balance for the solid phase is
0 = p

0 = (1 )ps +

vl (1 )s gk,
k

(7.276)

where ps is the pressure in the solid. For a granular solid, we can expect grain
motion to occur if the interactive force is large enough to overcome friction and
cohesion; the typical kind of criterion is that the shear stress satisfies
c + pe tan ,

(7.277)

but in view of the large confining pressure and the necessity of dilatancy for soil
deformation, the piping criterion will in practice be satisfied at the toe of the dam
(i.e. the front), and piping channels will eat their way back into the dam, in much
the same way that river drainage channels eat their way into a hillslope. A simpler
criterion at the toe then follows from the necessity that the effective pressure on the
grains be positive. A lucid discussion by Bear and Bachmat (1990, p. 153) indicates
that the solid pressure is related to the effective pressure pe which controls grain
deformation by
pe = (1 )(ps p),

(7.278)

and in this case the piping criterion at the toe is that pe < 0 in the soil there, or
pe /z > 0. From (7.275), (7.276) and (7.278), this implies piping if
v
> (s l )(1 )g,
k

(7.279)

where v is the vertical component of vl . This criterion is given by Bear (1972).


More generally, piping can be expected to occur if pe reaches 0 in the soil interior
(ignoring cohesion). Sellmeijer and Koenders (1991) develop a model for piping.

7.12

Notes and References

451

Taylor Dispersion Taylor dispersion is named after its investigation by Taylor


(1953), who carried out experiments on the dispersal of solute in flow down a tube.
The dispersion is enabled by the combination of differential axial advection by the
down tube velocity, typically a Poiseuille flow, and the rapid cross stream diffusion
which renders the cross-sectional concentration profile radially uniform. The theory
of Taylor is somewhat heuristic; it was later elaborated by Aris (1956). For a formal
derivation using asymptotic methods, see Fowler (1997, p. 222, Exercise 2).
Its application to porous media stems from the conceptual idea that the pore space
consists of a network of narrow tubules connected at pore junctions. If the tubes are
of radius a and length dp , the latter corresponding
to grain size, then the Darcy
flux |u| U , while the pore radius a dp . This would suggest a Taylor
dispersion coefficient of
DT

dp2 |u|2
a2U 2
,

48D
48 2 D

(7.280)

as opposed to the measured values which more nearly have DT |u|. Taylor dispersion in porous media has been studied by Saffman (1959), Brenner (1980) and
Rubinstein and Mauri (1986), the latter using the method of homogenisation.
Biofilm Growth Monod kinetics was described by Monod (1949), by way of
analogy with enzyme kinetics, where one considers the uptake of nutrients as occurring through a series of fast intermediary reactions; when two nutrients control
growth, as in respiration, it is usual to take the growth rate as proportional to the
product of two Monod factors (Bader 1978). A variety of enhancements to this
simple model have also been proposed to account for nutrient consumption due to
maintenance, inactivation of cells in adverse conditions, and other observed effects
(Beeftink et al. 1990; Wanner et al. 2006).
Bacteria in soils commonly grow as attached biofilms on soil grains, with a thickness of the order of 100 . A variety of models to describe biofilm growth have been
presented, with an ultimate view of being able to parameterise the uptake rate of
contaminant species in soils and other environments (Rittmann and McCarty 1980;
Picioreanu et al. 1998; Eberl et al. 2001; Dockery and Klapper 2001; Cogan and
Keener 2004).
Remediation Sites The three sites described in Sect. 7.8 are under study by the
Groundwater Restoration and Protection Group at the University of Sheffield, led by
Professor David Lerner. The site at Four Ashes is described by Mayer et al. (2001),
that at Rexco by Httmann et al. (2003), and that at St. Albans by Wealthall et al.
(2001).
The description of the two species reaction front given by (7.192) is similar to
that for a diffusion flame (Buckmaster and Ludford 1982) in combustion, and also
corrosion in alloys (Hagan et al. 1986). It is not conceptually difficult to extend
this approach to an arbitrary number of reactions, although it may become awkward
when multiple reaction fronts are present (see, for example, Dewynne et al. 1993).

452

7 Groundwater Flow

Diagenesis The first order reaction kinetics (7.225) for the smectiteillite transition was proposed by Eberl and Hower (1976). Information on solubility limits is
given by Aagaard and Helgeson (1983) and Sass et al. (1987). The asymptotic approximation called here the weak solubility limit is called solid density asymptotics
by Ortoleva (1994). Details of the use of the weak solubility approximation can be
found in Fowler and Yang (2003).
Compaction Interest in compaction is motivated by its occurrence in sedimentary
basins, and also by issues of subsidence due to groundwater or natural gas extraction
(see, for example, Ba et al. 2000). The constitutive law used here for effective
pressure is that of Smith (1971); it mimics the normal consolidation behaviour of
compacting sediments (such as soils), and is further discussed by Audet and Fowler
(1992) and Jones (1994).
Athys law comes from the paper by Athy (1930). Smith (1971) advocates the use
of the high exponent m = 8 in (7.265). Further details of the asymptotic solution of
the compaction profiles are given by Fowler and Yang (1998). Freed and Peacor
(1989) show examples of the flattened porosity profiles at depth.
Early work on pressure solution in sedimentary basins was by Angevine and Turcotte (1983) and Birchwood and Turcotte (1994). More recently, Fowler and Yang
(1999) derived the viscous compaction law. An extension to viscoelastic compaction
has been studied by Yang (2000).
Seals One process which we have not described is the formation of high pressure
seals. In certain circumstances, pore pressures undergo fairly rapid jumps across
a seal, typically at depths of 3000 m. Such jumps cannot be predicted within the
confines of a simple compaction theory, and require a mechanism for pore-blocking.
Mineralisation is one such mechanism, as some seals are found to be mineralised
with calcite and silica (Hunt 1990). In fact, a generalisation of the clay diagenesis
model to allow for calcite precipitation could be used for this purpose. As it stands,
(7.232) predicts a source for , but mineralisation would cause a corresponding sink
term. Reduction of leads to reduction of diffusive transport, and the feedback is
self-promoting. Problems of this type have been studied by Ortoleva (1994), for
example.

7.13 Exercises
7.1 Show that for a porous medium idealised as a cubical network of tubes, the permeability is given (approximately) by k = dp2 2 /72 , where dp is the grain
size. How is the result modified if the pore space is taken to consist of planar sheets between identical cubical blocks? (The volume flux per unit width
between two parallel plates a distance h apart is h3 p  /12, where p is the
pressure gradient.)

7.13

Exercises

453

7.2 A sedimentary rock sequence consists of two types of rock with permeabilities
k1 and k2 . Show that in a unit with two horizontal layers of thickness d1 and
d2 , the effective horizontal permeability (parallel to the bedding plane) is
k = k1 f1 + k2 f2 ,
where fi = di /(d1 + d2 ), whereas the effective vertical permeability is given
by
1
k
= f1 k11 + f2 k21 .

Show how to generalise this result to a sequence of n layers of thickness


d 1 , . . . , dn .
Hence show that the effective permeabilities of a thick stratigraphic sequence containing a distribution of (thin) layers, with the proportion of layers
having permeabilities in (k, k + dk) being f (k) dk, are given by


f (k) dk
1
kf (k) dk,
k
=
.
k =
k
0
0
7.3 Groundwater flows between an impermeable basement at z = hb (x, y, t) and
a phreatic surface at z = zp (x, y, t). Write down the equations governing the
flow, and by using the Dupuit approximation, show that the saturated depth h
satisfies
kg
ht =
.[hzp ],

where = (/x, /y). Deduce that a suitable time scale for flows in an
aquifer of typical depth h0 and extent l is tgw = l 2 /kgh0 .
I live a kilometer from the river, on top of a layer of sediments 100 m thick
(below which is impermeable basement). What sort of sediments would those
need to be if the river responds to rainfall at my house within a day; within a
year?
7.4 A two-dimensional earth dam with vertical sides at x = 0 and x = l has a
reservoir on one side (x < 0) where the water depth is h0 , and horizontal dry
land on the other side, in x > l. The dam is underlain by an impermeable
basement at z = 0.
Write down the equations describing the saturated groundwater flow, and
show that they can be written in the dimensionless form
u = px ,

2 w = (pz + 1),

pzz + 2 pxx = 0,
and define the parameter . Write down suitable boundary conditions on the
impermeable basement, and on the phreatic surface z = h(x, t).
Assuming  1, derive the DupuitForchheimer approximation for h,
ht = (hhx )x

in 0 < x < 1.

454

7 Groundwater Flow

Show that a suitable boundary condition for h at x = 0 (the dam end) is


h = 1 at x = 0.
Now define the quantity


U=

p dz,
0

and show that the horizontal flux



q=

u dz =

U
.
x

Hence show that the conditions of hydrostatic pressure at x = 0 and constant (atmospheric) pressure at x = 1 (the seepage face) imply that
 1
1
q dx = .
2
0
Deduce that, if the Dupuit approximation for the flux is valid all the way to
the toe of the dam at x = 1, then h = 0 at x = 1, and show that in the steady
state, the (dimensional) discharge at the seepage face is
qD =

kgh20
.
2l

Supposing the above description of the solution away from the toe to be
valid, show that a possible boundary layer structure near x = 1 can be described by writing
x = 1 2 X,

h = H,

z = Z,

p = P ,

and write down the resulting leading order boundary value problem for P .
7.5 I get my water supply from a well in my garden. The well is of depth h0
(relative to the height of the water table a large distance away) and radius r0 .
Show that the Dupuit approximation for the water table height h is


h
h kg 1
=
rh
.

t
r r
r
If my well is supplied from a reservoir at r = l, where h = h0 , and I withdraw
a constant water flux q0 , find a steady solution for h, and deduce that my well
will run dry if
q0 >

kgh20
.
ln[l/r0 ]

Use plausible values to estimate the maximum yield (litres per day) I can use
if my well is drilled through sand, silt or clay, respectively.

7.13

Exercises

455

7.6 A volume V of effluent is released into the ground at a point (r = 0) at time t.


Use the Dupuit approximation to motivate the model


h
h kg 1
=
rh
,

t
r r
r
h = h0 at t = 0, r > 0,

r(h h0 ) dr = V /2,

t > 0,

where h0 is the initial height of the water table above an impermeable basement. Find suitable similarity solutions in the two cases (i) h0 = 0, (ii) h0 > 0,
h h0  h0 , and comment on the differences you find.
7.7 Fluid flows through a porous medium in the x direction at a linear velocity U .
At t = 0, a contaminant of concentration c0 is introduced at x = 0. If the longitudinal dispersivity of the medium is D, write down the equation which determines the concentration c in x > 0, together with suitable initial and boundary
conditions. Hence show that c is given by







1
x Ut
x + Ut
Ux
c
= erfc
erfc
+ exp
,
c0 2
D
2 Dt
2 Dt
where
2
erfc =

es ds.
2

[Hint: you might try Laplace transforms, or else simply verify the result.]
2
Show that for large , erfc = e [ 1 + ], and deduce that if x =

U t + 2 Dt , with = O(1), then


 
1
1
c
erfc + O .
c0 2
t
Hence show that at a fixed station x = X far downstream, the measured profile
is approximately given by

  3 1/2 

1
X
1 U
c c0 1 erfc
t
.
2
2 DX
U
This is called the breakthrough curve, and indicates that dispersion causes
breakthrough to occur over a time interval (at large distance) of order tb =
(DX/U 3 )1/2 . If D aU , show that the ratio of tb to tb = X/U is tb /tb
(a/X)1/2 .
7.8 Rain falls steadily at a rate q (volume per unit area per unit time) on a soil
of saturated hydraulic conductivity K0 (= k0 w g/, where k0 is the saturated

456

7 Groundwater Flow

permeability). By plotting the relative permeability krw and suction characteristic /d as functions of S (assuming a residual liquid saturation S0 ), show
that a reasonable form to choose for krw () is krw = ec . If the water table
is at depth h, show that, in a steady state, is given as a function of the dimensionless depth z = z/zc , where zc = /w gd ( is the surface tension, d
the grain size), by
 sinh 1 (ln 1 c)
1
1
2
q
h z = ln
,
1

2
c
sinh 2 ln q1

where h = h/zc , providing q = q/K0 < 1. Deduce that if h zc , then


1
1
c ln q near the surface. What happens if q > K0 ?
7.9 Derive the Richards equation



pc
S
k0
=
krw (S)
+ w g
w
t
z
z
for one-dimensional infiltration of water into a dry soil, explaining the meaning of the terms, and giving suitable boundary conditions when the surface flux
q is prescribed. Show that if the surface flux is large compared with k0 w g/,
where k0 is the saturated permeability, then the Richards equation can be approximated, in suitable non-dimensional form, by a nonlinear diffusion equation of the form


S

S
=
D
.
t
z
z
Show that, if D = S m , a similarity solution exists in the form
S = t F (),
where =

1
m+2 ,

m+1
m+2 ,

(F m F  ) = F F  ,

= z/t ,

and F satisfies
F m F  = 1 at = 0,

Deduce that
F m F  = ( + )

F 0 as .

F d F,

where 0 (which may be ) is where F first reaches zero. Deduce that F  < 0,
and hence that 0 must be finite, and is determined by
 0
1
.
F d =
+
0
What happens for t > F (0)1/ ?

7.13

Exercises

457

7.10 Write down the equations describing one-dimensional consolidation of wet


sediments in terms of the variables , v s , v l , p, pe , these being the porosity,
solid and liquid (linear) velocities, and the pore and effective pressures. Neglect the effect of gravity.
Saturated sediments of depth h lie on a rigid but permeable (to water) basement, through which a water flux W is removed. Show that
vs =

k p
W,
z

and deduce that satisfies the equation






k p

=
(1 )
W .
t
z
z
If the sediments are overlain by water, so that p = constant (take p = 0) at
z = h, and if = 0 + p/K, where the compressibility K is large (so 0 ),
show that a suitable reduction of the model is
p
2p
p
W
=c 2,
t
z
z
where c = K(10 )k/, and p = 0 on z = h, pz = W/k. Non-dimensionalise
the model using the length scale h, time scale h2 /c, and pressure scale
W h/k. Hence describe the solution if the parameter = W h/k is small,
and find the rate of surface subsidence. What has this to do with Venice?
7.11 Write down a model for vertical flow of two immiscible fluids in a porous
medium. Deduce that the saturation S of the wetting phase satisfies the equation





S
pc
q

+ g =
+
Meff
Meff
,
t
z
Mnw
z
z
where z is a coordinate pointing downwards,
pc = pnw pw ,

 = w nw ,



1
1
Meff
= Mw1 + Mnw
,

q is the total downward flux, and the suffixes w and nw refer to the wetting
and non-wetting fluid, respectively. Define the phase mobilities Mi . Give a
criterion on the capillary suction pc which allows the BuckleyLeverett approximation to be made, and show that for q = 0 and w nw , waves typically propagate downwards and form shocks. What happens if q = 0? Is the
BuckleyLeverett approximation realistice.g. for air and water in soil? (Assume pc 2 /rp , where = 70 mN m1 , and rp is the pore radius: for clay,
silt and sand, take rp = 1 , 10 , 100 , respectively.)
7.12 A model for snow-melt run-off is given by the following equations:


k pc
u=
+ l g ,
z

458

7 Groundwater Flow

k = k0 S 3 ,
S u
+
= 0,
t
z


1
S .
pc = p0
S

Explain the meaning of the terms in these equations, and describe the assumptions of the model.
The intrinsic permeability k0 is given by
k0 = 0.077 d 2 exp[7.8s /l ],
where s and l are snow and water densities, and d is grain size. Take d =
1 mm, s = 300 kg m3 , l = 103 kg m3 , p0 = 1 kPa, = 0.4, = 1.8
103 Pa s, g = 10 m s2 , and derive a non-dimensional model for melting
of a one metre thick snow pack at a rate (i.e. u at the top surface z = 0) of
106 m s1 . Determine whether capillary effects are small; describe the nature
of the model equation, and find an approximate solution for the melting of an
initially dry snowpack. What is the (meltwater flux) run-off curve?
7.13 Consider the following model, which represents the release of a unit quantity of groundwater at t = 0 in an aquifer < x < , when the Dupuit
approximation is used:
ht = (hhx )x ,
h = 0 at t = 0, x = 0,

h dx = 1

(i.e., h = (x) at t = 0). Show that a similarity solution to this problem exists
in the form
h = t 1/3 g( ),

= x/t 1/3 ,

and find the equation and boundary conditions satisfied by g. Show that the
water body spreads at a finite rate, and calculate what this is.
Formulate the equivalent problem in three dimensions, and write down the
equation satisfied by the similarity form of the solution, assuming cylindrical
symmetry. Does this solution have the same properties as the one-dimensional
solution?
7.14 The tensor Dij (i, j = 1, 2, 3) has three invariants
DI = Dii ,

DI I = Dij Dij ,

DI I I = Dij Dj k Dki .

(Summation over repeated indices is implied.) Show that the invariants of the
tensor
ui uj
,
Dij = uij + ( )
u

7.13

Exercises

459

where u = |u| and ij is the Kronecker delta (= 1 if i = j , = 0 if i = j ), are


the same as those of the tensor

0
0
 u
u
0 .
D= 0
0
0
u
7.15 Suppose that a doubly porous medium consists of a periodic sequence of
blocks M with boundaries (fractures) M. The concentration of a chemical
reactant c is taken to be a function of the fast space variable X and the slow
space variable x = X, and we assume that c = c(x)

+ 2 ck (X), where the


suffix k refers to fractures (f ) or matrix block (m).
Let G(X, Y) be a Greens function satisfying
Y2 G = (X Y) in M,

G = 0 for Y M,

and suppose that


X2 = 0 in M,
= on M.
Show that


=
M

where

G
NY

G(X, Y)
(Y) dS(Y),
NY

= n. Y G(X, Y), and hence show that





=
K(X, Y)(Y) dS(Y),
N M
M

where
K(X, Y) =

2 G(X, Y)
.
NX NY

Now suppose that the fluctuating matrix and fracture concentrations of the
chemical reactant are given by


c
X2 cm = Pe
+ um .x c + Pec x2 c Rm in M,
t
subject to cm = cf on M, and

(1 f ) cm 

f
N M


c
+ uf . x c + Pec x2 c Rf
= Pe
t

X2 cf

subject to conditions of periodicity with zero mean.

on M,

460

7 Groundwater Flow

Show that, if we define c to be the solution of


X2 c = 1 in M,
with
c = 0 on M,
then


cm =
M

G(X, Y)
cf (Y) dS(Y) + Rm c ,
NY

and deduce that, for X M,

f X2 cf (1 f )

K(X, Y)cf (Y) dS(Y)


M


c 
= f Rf + (1 f )Rm
.
N M
By integrating this equation over M, show that the condition of periodicity
of cf implies that the equation to determine c is


c

+ u. x c = x2 c Pec,
Pe
t
where u = f uf + (1 f )um .
7.16 The reaction rates in the reactions
r1

SS XL + nH2 O,
r2

L
KFs K+L + AlOL
2 + s SiO2 ,
r3

L
S
L
K+L + AlOL
2 + f X f I + SiO2 ,

SiOL
2

r4+

 Qz,

r4

are related by
r 1 f r3 ,
r2 r 3 ,
r4+

r4 (s + 1)r3 .

The reaction rate r3 is given by




f r3
r3
r3 = I R3 1
1

S R1
F R2 L 1


(s+1)r3 
Q R4+

7.13

Exercises

461

where i are porosities, Rk are rate factors (such that rk Rk ), and the stoichiometric constants f and s, and the constant L , may be taken as O(1) (and
L > 1). Show that r3 can be written explicitly in the form


2
1
s +1
f
1
=
+
+
+
r3
(L 1)F
(L 1)Q S
I


s +1
f
1 2
1
+
+
+
+
(L 1)F
(L 1)Q S
I

 1/2
f
4(s + 1)
1
+
+
,
(L 1)Q S I
where the coefficients Y represent the porosity weighted rate factors, i.e.,
I = I R3 ,

S = S R1 ,

Q = Q R4+ ,

F = F R2 .

Deduce that the slowest reaction of the four (as measured by Y ) controls the
overall rate, and give explicit approximations for r3 for each of the consequent
four possibilities.

You might also like