You are on page 1of 43

FACULTY OF ENGINEERING, TECHNOLOGY, APPLIED

DESIGN AND FINE ART

DEPARTMENT OF CIVIL ENGINEERING

HYDRAULICS AND HYDROLOGY

LECTURE 4
4.0 INFILTRATION AND SOIL MOISTURE

4.1 Infiltration

When rain falls upon the ground it first of all wets the vegetation or the bare soil. When
the surface cover is completely wet, subsequent rain must either penetrate (infiltrate) the
surface layers if the surface is permeable, or run off the surface towards a steam
channel if the surface is impermeable.
Infiltration therefore refers to the entry of water into the soil, a combined force process of
capillary attraction and gravitation along with pressure due to water ponding at the
ground surface. The infiltration rate is high at the initial stage as soil voids such as
animal burrow, root tunnels, macro and micro pores gradually fill with water. It will
eventually decline to a constant at which the infiltration rate is equal to the rate of water
drained out through the soil profile by gravity. This process of draining water to deeper
layers is called Percolation.
Interest in infiltration processes was first stimulated by RE Horton’s theory of runoff,
which identified its crucial role in dividing rainfall between storm runoff and ground water
seepage. According to Horton’s theory, the hydrologically Effective rainfall is all the
water that remains on the surface after infiltration and any losses in surface detentions,
interception and evaporation (See fig 4.1). Infiltration is clearly of great importance in the
global hydrological cycle, estimated at 76% of the world’s land-area precipitation
(L’vovich,1974). It is the process that provides all the water used by natural and
cultivated plants and almost all the water that enters the ground-water reservoirs.
Thus, understanding the linked processes of infiltration and redistribution (movement of
infiltrated water in the unsaturated zone of a soil) is an essential basis for many crucial
aspects of water-resource management, including;
(a) Developing strategies for crop irrigation
(b) Understanding chemical processes in soils, including natural weathering, and the
movement of natural nutrients, fertilizers, and pesticides; and
(c) Estimating the timing and amounts of ground water recharge.
In addition, much of the motivation for studying infiltration has come from the need to
forecast or predict hydrologic response to rain or snowmelt events. Water that does not
infiltrate typically moves relatively quickly as overland flow towards a stream channel
and contributes to short term stream response, perhaps causing flooding. In contrast,
infiltrated water is either retained in the soil and ultimately evapotranspired or moves,
usually relatively slowly, to the surface-water system through underground paths.

2
Infiltration rate

Rainfall

Surface Runoff
Rainfall and fo
infiltration
-1
rates (mmhr ) Infiltration capacity
Surface Detention

Infiltrated Water fc

Time (hours)
Fig 4.1: The role of saturated infiltration capacity and Horton’s
“effective rainfall” in generating surface runoff

4.1.1 Infiltration Rate and Infiltration Capacity

An infiltration rate is the volumetric rate of seepage into the upper surface of the soil,
expressed as mm3 per mm2 of surface (i.e mm) over a period of time, e.g. mm s-1. Rates
are highly variable, depending on the soil and its current water content; typically higher
in a dry soil and reducing during rainfall. In contrast, the infiltration capacity is more
stable and may be treated as a constant parameter of a given soil type. It is the
maximum rate at which a given soil type can absorb water when it is in a specified
condition. The most meaningful definition is the saturated infiltration capacity
which Horton (1937) termed the limiting infiltration capacity, ‘the relatively low and
steady infiltration rate in a soil surface free from cracks and major holes, that has
been wetted to saturation point for long enough to permit full swelling of clays and
colloids and full adjustment of soil structure to a stable saturated state’. The aim of
this definition is to produce an easily measurable, stable and reproducible value, while
making simplifying assumptions (in this case, about cracks) that may limit its
applicability in some circumstances.
Saturated infiltration capacities vary with soil properties, vegetation, soil faunal activity,
land management, slope, geomorphic location and temperature.

The infiltration takes place at capacity rates only when the intensity of rainfall equals or
exceeds fc; i.e, f = f when i ≥ f ; but when i < f , f < f and the actual infiltration
c c c c

rates are approximately equal to the rainfall rates. The infiltration depends upon
the intensity and duration of the rainfall,, weather (temperature), soil

3
characteristics, vegetation cover, land-use, initial soil moisture content,
entrapped air and depth of the ground water table.

4.1.2 Factors affecting Infiltration Rate

Water movement into and through soil profiles is affected by a variety of factors
reflecting the surface and subsurface conditions and flow characteristics.
A) Surface conditions include;
(i) Vegetation cover – Vegetation can increase infiltration by creating cracks
around the stem or trunk, and root holes, or simply by protecting the soil
surface from rain splash and sun-baking. Even infilled dead stump holes can
still have higher infiltration capacity (Brewer and Sleeman, 1963). By increasing
surface roughness and holding up surface water drainage, vegetation and
forest litter also increase the total amount of infiltration and by transpiring and
drying the soil it will tend to increase initial infiltration rates. For a given soil, the
infiltration in a vegetated surface e.g forest can be many times greater than that
over bare ground. Vegetation cover also provides protection against rain drop
impact and helps to increase infiltration.
(ii) Land management practices – Both soil and vegetation can be sensitive to land
management practices. Different crops and agricultural practices can create
greater variability in infiltration capacities than differing natural vegetation.
Ploughing is intended to improve infiltration and drainage although some soils
may develop impermeable plough-pans.
(iii) Crusting – Heavy machinery, tramping animals, and over exposure of bare
soils to heavy rain cause compacted surfaces hence reduced infiltration
capacities. Mulching can reduce soil compaction in arable fields.
(iv) Cracking – Soils with massive structure, with no cracks or substantial voids,
are least porous, whereas blocky or prismatic structure favours infiltration along
vertical cracks or pores.
(v) Slope and topographic location affect both surface and subsurface drainage.
However, these effects tend to get confounded with those of soil type, since
soil drainage is an important factor in the development of different soil types.
For example, it affects chemical processes by controlling aeration as well as
the eluviation and deposition of finer mineral and organic material. For a
given slope angle, infiltration rates are likely to be higher on convex or
‘shedding’ slopes than in concave or ‘receiving’ sites, because of better soil
drainage at shedding sites. This is due partly to higher hydraulic gradients
within the soil and partly to coarser grained soils in shedding sites.
(vi) Surface temperatures – Low PH and Low temperatures can result in peat
accumulation. Infiltration measurements on peat generally indicate very low
capacity, but, like clay, peat expands and contracts quite noticeably during
wetting and drying, causing cracking. In the presence of substantial cracks,
most water is likely to enter via the cracks.
(vii) Chemicals – The acidity or base status of the soil can also affect infiltration
rates. Neutral to moderate alkaline soils tend to have the best crumb
structure, as a result of the biding action of calcium and magnesium. The
better drainage and nutrient status in these soils also attract soil fauna and
micro-organisms that create and maintain open pores and can excrete
colloids that bind the soil particles.

4
(viii) Soil erosion which exposes either the less permeable or the more
permeable soil from beneath the ground surface.
B) Conditions under the ground include; soil texture, structure, organic matter
content, depth, compaction, voids, layering, water content, ground water table and
root system. These factors affect the soil water-holding capacity and ability of water
to move.
4.1.3 Variation of Infiltration Capacity with Time

Figure 4.2 shows the relationship between the infiltration capacity of soil and time. It
represents the instantaneous rate of water (cm/hr) entering the soil medium, after the
start of rainfall or application of water with respect to time lapse. The intake rate declines
with time.

Infiltration
Sandy soil
Capacity

(cm/hr) Moderately textured soil


Fine textured soil

Time, h

Fig 4.2: Relationship between the intake rate of soil and time

4.1.4 Variation of Total Intake with Time

The relationship between the total intake of soil with time (Fig. 4.3) is in the form of an
expression of the time required for a soil to absorb a specified amount of water. It helps
to determine the speed at which the wet front advances in a field.

Sandy soil Moderately textured soil

Intake
Fine textured soil
(cm)

Time, h

Fig 4.3: Relationship between the total Intake rate of soil and time

5
4.2 Methods of Determining Infiltration

The methods of determining infiltration include;

 Infiltrometers
 Observation in Pits and Ponds
 Placing a Catch basin below a laboratory sample
 Artificial rain simulators
 Hydrograph analysis

4.2.1 Infiltrometers

In this method, characteristics of a soil are determined by ponding the water in a metal
cylinder after installing it on the ground surface and then measuring the rate of
absorption of ponded water into the ground surface by registering the fall of water level
in the cylinder.
An infiltrometer is a wide diameter, short tube surrounding an area of soil. A single tube
infiltrometer may be used but this gives a high degree of variability in the measured
data due to uncontrolled lateral movement of water from the cylinder after the wetting
front reaches the cylinder bottom. For this reason, two rings (Double ring
infiltrometers) are usually used to minimise to some extent the edge effects of the
surrounding drier soil and to prevent water within the inner space from spreading over a
larger area after penetrating below the bottom of the ring. This is achieved by ponding
water in buffer area around the first cylinder by providing a concentric cylinder of larger
diameter, which also acts as a guard cylinder. The measurement by cylinder
infiltrometers is affected by the thickness of the cylinder material, the type of bevelling of
the cylinder bottom, the method of driving the cylinder into the soil and the depth of
cylinder installation.
Method of Measurement
Two concentric cylinders of rolled steel 2mm thick, 25cm height and off diameters 30mm
and 50mm are driven into the ground up to a depth of 15cm with a wooden mallet. A
point gauge is fixed on the inner cylinder for measurement of the depth of water in it as
shown in Fig. 4.4.

6 to 12 cm
GSL
30 cm 10 cm
15cm

Fig 4.4: Concentric ring infiltrometers installed in the field.

A measured quantity of water is added to the inner cylinder up to the desired level. To
prevent any puddling or sealing of the ground surface inside the inner cylinder, a jute
matting is recommended to be placed while pouring water. After the cylinder is filled to
three-quarters of the height from the ground surface, the matting is removed. The buffer
area between the two cylinders is also filled with water and the level of water is
maintained approximately at the same level in both cylinders. The water depth is
measured at regular intervals and infiltration is calculated for each interval. Care is taken

6
to add water after each reading in order to maintain a constant head in the inner
cylinder, for infiltration. Samples of the recorded results are shown in Table 4.1.
Table 4.1 Results from Infiltrometer Measurements
Elapsed time Water surface distance from a Infiltration during elapsed time
(min) Reference level
Before filling After filling Depth Average Accumulated
(cm) (cm) (cm) (cm/h) Depth (cm)
0 0 10.0 - - -
5 8.3 10.0 1.7 20.4 1.7
10 9.0 10.0 1.0 12.0 2.7

The infiltration rate of the soil for an area is computed by taking measurements at
different points of the soil and then averaging the data collected. Care should be taken to
avoid points where there are abnormalities in the soil like soil crackling plant roots
surface sealing etc in the soil
Such tests give useful comparative results but they do not simulate real conditions and
have been largely replaced by sprinkler test of larger areas. Here the sprinkler simulates
rainfall, and runoff from the plot is collected and measured as well as inflow. The
difference assumed to have infiltrated.
Though rain simulating sprinklers are good deal more realistic than flooded rings, there
are still limitations to the reliability of results thus obtained, which give higher values of
infiltration than natural condition do.

4.2.2 Observation from infiltration pits and ponds

By noting the depression in the level of water in the pits and ponds and deducting the
loss due to evaporation, an idea about the infiltration rates in such soils can be obtained
(Raghunath, 2006).
4.2.3 Artificial Rain Simulators
On a small area of land of 0.1 to 50m2, water is applied by artificial showers at a uniform
rate. The resulting surface runoff is measured and the infiltration capacity of the soil is
determined (Raghunath, 2006).

4.2.4 Using Lysimeters


In this method, by placing a catch basin called a lysimeter under a laboratory sample or
at some depth below the land surface, the infiltrating water can be measured and
infiltrated rate in the soil can be obtained.

4.2.5 Hydrograph Analysis


By knowing accurately the varying intensities of rainfall during a storm and the
continuous record of the resulting runoff, the infiltration capacity can be determined. This
will be discussed further in the next course.

7
4.3 Horton’s Infiltration Equation

Horton proposed an entirely empirical infiltration equation to represent the generally


observed decrease of infiltration rate with time, tending to a steady-state value.
For water supply rate in excess of infiltration capacity,

(
f = f c + ( f 0 − f c ) e − kt 0 ≤ t ≥ t d ) (4.1)
fo − fc
k=
Fc
f = Infiltration rate of at anytime t (mm/hr) from start of the rainfall
fc = Final constant rate of infiltration (mm/hr) at saturation - saturated infiltration capacity
fo = Initial rate of infiltration capacity at t =0 (mm/hr)
t = Time from the start of the rainfall (min)
td = Total duration of rainfall (min)
F = shaded area in figure 4.5.
k = Constant depending on a particular soil and vegetation cover (min-1)

Figure 4.5: Infiltration curve (Horton)

Determination of fo and fc is tedious and difficult, therefore the above equation is not
popular and is thus rarely used.
The method has been widely used by hydrologists but in practise the equation tends to
underestimate the usually very rapid initial decrease of infiltration rate with time. An
example (4.1) is given.

8
Example 4.1
The infiltration capacities of an area at different intervals of time are indicated in Table
4.2. Find the equation for the infiltration capacity in the exponential form.
Table 4.2 Infiltration Capacities
Time in hours 0 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00
Infiltration capacity, f
(cm/h) 11.4 6.2 3.5 2.3 1.7 1.3 1.2 1.1 1.1

Solution
Using the general curve equation, f = f c + ( f 0 − f c )e − Kt

1
This is also expressed as t= −1
K log10 e log10 ( f − f c ) + log10 ( f o − f c ) and the
K log e
separate infiltration parameters are plotted in Table 4.3
Table 4.3 Infiltration Parameters
T (h) 0 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00

f (cm/h) 11.4 6.2 3.5 2.3 1.7 1.3 1.2 1.1 1.1

f - fc 10.3 5.1 2.4 1.2 0.6 0.2 0.1 0.0 0.0

Log 10 (f – fc) 1.01 0.70 0.38 0.08 -0.22 -0.70 -1.00 - -

A graph of Time vs Log (f – f ) is plotted in Fig 4.6.

9
2.5

y = -0.7527x + 0.7833

1.5
Time (h)

0.5

0
-1.5 -1 -0.5 0 0.5 1 1.5
log10 (f - fc)

Fig 4.6 A Plot of Time vs Log

The slope of the line = -0.7525

From the equation −1 = - 0.7525 K = 0.49


K log10 e

Hence the equation for Infiltration capacity curve is given as

f = 1 + (11.4 – 1.1) e -0.49t


f = 1.1 + 10.3e -0.49t

10
4.4 Infiltration Indices
Infiltration indices are useful in estimation of runoff Volume from larges areas. Infiltration
indices assume a constant average infiltration rate during a storm, although in actual
practice the infiltration will be varying with time. This is also due to different states of
wetness of the soil after commencement of the rainfall. There are three type of ifiltration
indices;

(i) Phi-Index
(ii) W-Index
(iii) fave-index – In this method an average infiltration loss is assumed throughout the
storm, for the period i > f

4.4.1 Infiltration φ-Index

The ϕ–index, defined as that rate of rainfall above which the rainfall volume equals the
runoff volume. ϕ–index is the most commonly used method for approximation of
infiltration losses. The method assumes a constant value for the intake rate of rainwater
into the soil for the full duration of the rainfall. The ϕ–index is relatively simple and all
losses due to infiltration, interception, and depression storage (i.e in pits, ponds etc) are
accounted for; hence
Ba sin Re ch arg e
φ − index = ,
Duration of ra inf all
Provided i > φ throughout the storm.

The φ -index gives the average infiltration of rainwater (storm loss). The remaining
volume of rainfall, which flows out as surface runoff is called the excess rainfall volume
(net rain). The time period during which the excess rainfall occurs is called the effective
time period

11
4.4.2 Infiltration φ-Index
Another method known as W Index allows for depression storage, short rainless periods
during the storm and eliminates all rain periods during which i < f c but since
determination of the initial values of abstraction is difficult, this method is not popular.
The W-index is more accurate than the φ -index because it excludes the interception and
depression losses which are considered as a part of infiltration in the latter. Thus W-
index is always less than φ -index.

W-index is the average rate of infiltration during the period when the rainfall intensity
exceeds the infiltration capacity rate and is given by

W-Index = P- R - S
tf
where P = total rainfall, R = surface runoff, S = effective surface retention, t f = time
period during which the rainfall intensity exceeds infiltration rate i > f c ; sometimes, it is
taken equal to the rainfall period, and W = average rate of infiltration (cm/hr).

Wmin index, the minimum values of W-index is obtained when the soil is very wet. At that
stage, because the effect of the depression storage and interception losses is negligible,
W-index are approximately equal. It may be noted that both W-index and φ -index vary
from storm to storm even for the same basin.

12
Procedure for Determination of the ϕ–Index
i. Draw the hyetograph of the storm rainfall and compute the total volume of
rainfall
ii. Compute the excess runoff volume from the storm hydrograph by separating
the base flow from it
iii. Subtract (ii) from (i) to determine the total intake of rainwater into the soil
iv. Divide the value at (iii) with the effective rainfall period to get the ϕ–index
(cm/hr)
Therefore
Total int ake of water into the soil
ϕ − index(cm/ hr) =
Effectiverainf all period (tc )
Total inf iltration − Infiltration during period when no excess rain occurs
=
Effective rainf all period (tc )

Example 4.2

A storm with 13.1cm precipitation produced a direct runoff of 6.0cm. The time distribution
of the storm is as given in the Table 4.4. Estimate the φ-index of the storm. Also
determine the W-index.
Table 4.4 Time Distribution of the Storm
Time from 1 2 3 4 5 6 7 8
start of storm
Incremental 0.45 0.7 1.83 2.5 3.36 2.7 0.97 0.59
rainfall in
each hour

Solution
4.00

Rainfall 3.50
intensity, 3.00
cm/h
2.50

2.00

1.50

1.00
Phi-index
0.50
=1.1
0.00
1 2 3 4 5 6 7 8
Time, h

Total precipitation P = 13.1 cm


Total Runoff Q = 6.0 cm

13
P − Q 13.1 − 6
Windex = =
tr 8
= 0.89 cm / hr
Since the ϕ–index should be more than 0.89 (non-uniform rainfall). The rainfall in the first
second and eighth hour is ineffective in producing excess rain. Therefore the effective
time period, te = 8 – 3= 5 hrs.
Infiltration during effective ra inf all period
ϕ − index =
te
8x0.89 − 0.45 − 0.7 − 0.59 5.38
= =
5 5
= 1.076
Check
The ϕ–index is marked off on the hyetograph as shown above. The area of the hatched
area above the ϕ–index line should equal to the given runoff.
= (1.83 – 1.076) + (2.5 – 1.076) + (3.36 – 1.076) + (2.7 –1.076)
= 6 cm which is as given

Example 4.3
A catchment area of 25km2 has one recording gauge. During a storm, the mass curve of
the rainfall was recorded in Table 4.5 as follows:
Table 4.5 Mass Curve of Rainfall
Time from start of storm 0 3 6 9 12 15 18 21
Accumulated rainfall 0 10 21 39 50 65 79 100
If the volume of rainfall due to the storm is 1.5 x 10 6 m3, estimate the ϕ–index of the
catchment.
Solution
Total runoff volume = 1.5 x 106 m3
Catchment area = 25 km2
Total runoff volume
∴Q = Total runoff depth =
Catchment area
1.5 x 10 6 m 3
=
25 x 10 6 m 2
= 0.06m = 60mm
Total Precipitation, P = 100mm
Total Infiltration = P – Q = 100 – 60 = 40mm

14
P − Q 100 − 60 40
W index = = = mm / hr
tr 21 21
= 1.9048 mm / hr
Since the ϕ–index should be greater than the W-index, the ϕ–index is a little more than
1.9048mm/hr, therefore causing infiltration to the ground every three- hour interval, more
than 3 X 1.9048 = 5.714 mm.
Incremental rainfalls are worked out in the Table 4.6. From the values of incremental
rainfall calculated it can be observed that, no excess rainfall occurred in two out of seven
(those values less than 5.714- the average value of infiltration in three hours). Hence the
excess rain must have fallen in 5 x 3 = 15hours. Hence the effective time period, t e =
15hrs.
Table 4.6 Incremental Rainfall

Time from start of Accumulated Rainfall Incremental rainfall in each


storm (mm) interval
0 0 -
3 3 3*
6 8 5*
9 26 18
12 50 24
15 70 20
18 89 19
21 100 11
40 − 5 − 3
ϕ − index =
15
= 32
15
= 2 . 133 mm / hr ( W index = 1 . 9048 mm / h )

Example 4.4
The following mass curve shown in Table 4.7 was obtained for a 21 hour effective
rainfall period that occurred on the catchment area of 40km2.Calculate the ordinates of
the effective rainfall hyetograph and the runoff volume, when the ϕ– index is 0.5cm/hr.
Table 4.7 Mass Curve

Time from start of storm (h) 0 3 6 9 12 15 18 21


Accumulated rainfall 0 1.5 4.0 6.0 6.5 9.0 12.0 14.0

Solution
A tabular solution of the problem is given Table 4.8. First the incremental rainfall in a
time, t is calculated. The infiltration losses in a given interval are obtained by multiplying
the ϕ–index with the corresponding time interval. The hyetograph ordinates are then

15
obtained by subtracting the infiltration losses from the corresponding incremental rainfall
in a time t. The ordinates of the rainfall intensity are obtained by dividing by the time
interval.

Table 4.8 Determination of Infiltration Parameters

Time interval Accumulated Rainfall in Infiltration Hyetograph Rainfall


τ (h) Rainfall (cm) Time τ(cm) Losses –x τ Ordinate(cm) Intensity
(cm) (cm/hr)

0-0 0.0 0.0 0.0 0.0 0.00


0-3 1.5 1.5 1.5 0.0 0.00
3-6 4.0 2.5 1.5 1.0 0.33
6-9 6.0 3.5 1.5 2.0 0.67
9-12 6.5 3.0 1.5 1.5 0.50
12-15 9.0 6 1.5 4.5 1.50
15-18 12 6 1.5 4.5 1.50
18-21 14 8 1.5 6.5 2.17

Direct runoff volume from the catchment = summation of volumes of runoff during each
time period, from the catchment
= Area of the hyetograph bars x catchment area
Area of hyetograph bars = 3hr (0.33+0.67+0.50+1.5+1.5+2.17) cm/h
= 20.01cm =0.2001m
Catchments area = 40km2 =40x106m2
Therefore volume of direct runoff = (0.2001x40x106) m3
= (8.004x106) m3

16
4.5 Soil Moisture

4.5.1 Soil Structure and Composition


The structure and composition of the soil (Fig. 4.7) will, to a great extent determine the
amount of moisture a soil can hold. Thus it is important to appreciate the concepts of soil
structure and composition.
The original parent material of soil is the soil rock of the earth’s outer skin. Weathering
and erosion break down the surface layers of the soil geological strata and in many
areas large quantities of unconsolidated material (soil) have been deposited.

Fig 4.7 Soil composition

Thus soil may be a direct product of underlying weathered rocks or may be formed from
loose deposits unrelated to the rock below, but deposited by wind, water, ice etc. Soil
deposits and their composition can therefore be very variable. Another most important
constituent of a soil especially in the upper layers is the organic material derived from
decomposition of living plants and other organisms.

4.5.2 Material Properties of the Soil


4.5.2.1 Distribution of Pore and Particle sizes

The soil structure may be taken to consist of a matrix of solid grains (mineral and
organic) between which are interconnected pore spaces that can contain varying
proportions of water and air. To simplify the discussion, the organic component is often
ignored and the soil is considered as a three phase system consisting of solid mineral
grains, water and air. The volume of water vapour contained in the air in soil pores is
insignificant and can be neglected.

The size of the pores through which water flow occurs is approximately equal to the
grain size, and the distribution of the pore sizes is determined largely by the grain-size

17
distribution. Most soils are a mixture of grain sizes, and the grain size distribution is often
represented as a cumulative-frequency plot of grain diameter (logarithmic scale) versus
weight fraction of grains with smaller diameter. The steeper the slope
of such plots, the more uniform the soil grain-size distribution.

For many purposes, the particle size distribution is characterised by the soil texture,
which is determined by the proportions by weight of clay, silt, and sand.
shows the range of grain sizes in those classes. Different schemes have been
developed for defining textures e.g, the BS and U.S Department of Agriculture schemes.
According to the U.S Department of Agriculture scheme, the texture is determined by the
proportions of sand, silt and clay after particles larger than sand (i.e. >2mm) are
removed. If a significant proportion of the soil (>15%) is gravel or larger, an adjective
such as “gravelly” or “stony” is added to the soil texture term e,g a gravelly sandy loam.

Measurement: Weight fractions of soils of various diameters are measured by sieve


analysis for particles larger than 0.05mm (or 0.063mm) and by sedimentation for the
smaller grain sizes (see Hillel 1980).

4.5.2.2 Particle Density


Particle density, ρm, is the weighted average density of the mineral grains making up a
soil:
Mm
ρm = Where M m is the mass and Vm the volume of the mineral grains.
Vm
The value ρ m for a given soil is not usually measured, but is estimated based on the
mineral composition of the soil. A value of 2.65 g cm-3 , which is the density of the
mineral quartz, is assumed for most soils.

4.5.2.3 Bulk density

Bulk density, ρ b , is the dry density of the soil:


Mm Mm
ρb = = , Where Vs is the total volume of a soil sample and
Vs Va + Vw + Vm
Va , Vw and Vm are the volumes of the air, liquid water, and mineral components of the
soil, respectively. In most hydrologic problems, bulk density at any point is constant in
time; however, it commonly increases with depth due to compaction by the weight of
overlying soil.

In practice, bulk density is defined as the weight of a volume of soil that has been dried
for an extended period (16hours or longer) at 105oC divided by the original volume.
Typical values of bulk density (in gcm-3) are 0.7 for peats, 1.1 for clayey soils, 1.6 for
sands and 1.7 or more for naturally or artificially compacted loams. Thus, soils with more
spaces have lower bulk density.

18
4.5.2.4 Porosity

The sum of Va and Vw is the total volume of voids or pore spaces in the soil and Vs is the
total volume of a soil sample. The proportion of pore spaces in a volume of soil is the soil
porosity,Ø.

Va + Vw
φ=
Vs
Like bulk density, porosity is constant over the time periods considered in most
hydrologic analyses. However, in many cases, it decreases with depth due to
compaction and to the development of large pores by biological activity near the surface.

ρb
It can be shown that
φ = 1− Where φ is usually determined by measuring ρ b and
ρm
assuming an appropriate value for ρ m .

In general, finer-grained soils have smaller but more pore spaces and consequently
higher porosities, while coarser-grained soils have larger but less volume of spaces and
consequently lower porosities. Thus, the rate of water infiltration is higher, movement is
faster and water holding capacity is smaller for coarse soils. Peats which are highly
organic soils, may have porosities as high as 0.80.

4.5.3 Water in the Soil

Most of the content of a soil comes from rainfall or melting snow infiltrating as seepage
water moving by gravity and surface tension through the pore spaces. Its pathways are
smoothed by a thin film of hygroscopic water, held tightly by electrostatic forces on each
of the soil particles and is not easily moved by other forces including plant roots.
Below the percolating flow, the voids in the soil are filled with air and /or water vapour.
This layer is a zone of aeration with a complex mixture of solids, liquids and gases. With
increase in depth, the aeration zone gives way to a layer of saturated soil with all the
pore spaces occupied by water and the saturated capillary zone water is held by
capillary forces between the soil particles and is at a pressure less than atmospheric. At
greater depths in the same zone, the pressure exceeds atmospheric pressure. The
surface over which the pressure equals atmospheric pressure is defined as the water
table. The extent of the capillary zone is dependant on the soil composition and parking
of soil particles. We will discuss the soil water characteristic along a vertical profile in the
later section (s). The following definitions are key in understanding soil water
characteristics;
(i) Soil Water Content
Soil water can be volumetrically expressed by;

Vw
θv = . Thus, the volumetric water content, or simply water content, θ , is the ratio
Vs
of water volume to soil volume.

19
Water content may also be gravimetrically expressed by;
Mw
θw = , where M w is the mass of water content and M m is the mass of dry soil.
Mm
Converting θ w into θ v requires a multiplier, bulk density of the soil ρ b ; or

ρb
θ v =θw ( ) whereρ wis the density of water (1 gcm )
-3

ρw
In the laboratory, θ is determined by first weighing a representative soil sample of
known volume Vs , oven drying it at 105oC, re-weighing, and calculating

M swet − M sdry
θ= ; Where M swet and M sdry are the weights before and after drying,
ρ wVs
respectively, and ρ w is the density of water.

(ii) Degree of Saturation


The degree of saturation or wetness, S, is the proportion of pores that contain water.
Vw V θ
S= = w =
Va + Vw Vvoids φ
Wetness is not directly measured but calculated using the equation above.

4.5.3.1 Water in the Unsaturated Zone


In most areas, with the exception of bogs and swamps, the water table is some distance
below the ground surface. Between the ground surface and the water table is a region in
which the pore spaces of the rock or soil may be partly filled with air and partly with
water. This region is referred to as the unsaturated zone or vadose zone (“Vadose” is
from Latin word for “shallow”). Hydrologists want to be able to describe the flow of water
in the unsaturated zone to deal with a number of important issues. For example,
recharge of the ground water reservoir (aquifer). Recharge takes place most often
through the unsaturated zone, either overlying an unconfined aquifer or in the recharge
zone of a confined aquifer.

Another important aspect of water flow in the vadose zone is the water balance of plants.
Most terrestrial plants extract water from the vadose zone. Plants wilt when soils
become too dry because the forces holding the water in the soil are too great to allow
the plants access to the water. Related to the water balance of plants is the practice of
irrigation in agriculture. Agriculture accounts for about two-thirds of global water use,
with about 2.35x106 km2 of cropland irrigated worldwide, or about 16% of the total
(Postel, 1993). Arid regions are especially dependent on irrigation. One estimate of
world-wide average irrigation efficiency, the ratio of volume of water utilized by crops to
the total volume applied, is approximately 37% (Postel, 1993). Understanding the
movement of soil water, and its uptake by plants and "loss" through evapotranspiration
and recharge to the groundwater system, is essential in this regard.

20
The early literature recognized three divisions within the unsaturated zone: the capillary
fringe, the intermediate belt and the belt of soil water (e.g., see Meinzer, 1923).
According to Meinzer, the capillary fringe is "a zone in which the pressure is less than
atmospheric, overlying the zone of saturation and containing capillary interstices some
or all of which are filled with water that is continuous with the water in the zone of
saturation but is held above that zone by capillarity acting against gravity." That is, the
capillary fringe is a saturated zone above the water table where water is affected by
capillary forces. The uppermost belt, or belt of soil water, is "that part of the lithosphere
immediately below the surface, from which water is discharged into the atmosphere in
perceptible quantities by the action of plants or by soil evaporation." This definition
recognizes that plants, for the most part, extract water from a portion of the soil (the "root
zone") near the surface. The intermediate belt is "that part of a zone of aeration [i.e., the
unsaturated zone] that lies between the belt of soil water and the capillary fringe." The
intermediate belt is distinguished mainly by the fact that something must be between the
root zone and the capillary fringe.
Figure 4.8 shows the distribution of moisture in the vadose zone and classification of
waters according to Meinzer (1923). Water near the surface of the soil is available for
uptake by plant roots. After several days of fair weather, the moisture content in this belt
of soil water (or root zone) decreases substantially due to evapotranspiration. Directly
beneath the root zone, the moisture content tends to be fairly constant over a depth of
up to a meter or more. The relatively constant value of moisture content in this region is
referred to as the field capacity of the soil. Near the water table, the pores of the soil act
as "capillary tubes" and remain saturated even though the pressure head in the water is
negative. This saturated zone above the water table is the capillary fringe.
The ground water zone (some times called the phreatic zone – “Phreatic” comes from
Greek word for “well”) is saturated and the pressure is positive. If there is no ground
water flow, the pressure will be hydrostatic – that is, increasing linearly with depth
according to;
p( z ) = γ w ( z − z 0 ) γ
where p is gage pressure, w is the weight density of water, z is
distance measured vertically down-ward, and z0 is the value of z at the water table.
Thus the water table is at atmospheric pressure; it is the level at which water would
stand in a well.

21
Fig 4.8: The distribution of moisture in the vadose zone
and classification of waters according to Meinzer (1923)

4.5.3.2 Forces on Water in the Unsaturated Zone


The unsaturated zone is a three-phase system consisting of soil, water, and air. The
physical description of the system and of the flow of water in the system are thus more
complex than for the two-phase system of the saturated zone. A full treatment on a
microscale of the diverse forces acting on water in an unsaturated soil and the resultant
motion of this water is not feasible. Fortunately, empirical work shows that, as in
saturated soils, water flow in unsaturated soils is down a gradient of hydraulic head of
soil water. For flow of groundwater in the saturated zone, the hydraulic head is
composed of two components, pressure head (p/ρg) and head due to gravity, or
elevation head (z). Darcy's law states that the flux of groundwater is proportional to the
gradient in hydraulic head. Gradients in elevation head and in pressure head also drive
flows in the vadose zone. All terrestrial water, including soil moisture, is within the
Earth's gravitational field. Therefore, head due to gravity in the vadose zone is once
again the potential energy per unit weight, z, that is, the elevation above datum. The
main way that the physics for the vadose zone differs from that for the saturated zone is
in the pressure head term of the force balance.

4.5.3.2.1 Pressure head

If water is withdrawn from a rock or soil matrix that does not shrink upon drying, air
enters the pore space, and air-water interfaces (menisci) are present in the pore space.
Such curved interfaces are maintained by capillary forces. Surface tension acting in the
interfaces provides a mechanism of soil-water retention against externally applied
suction. This phenomenon is seen in the rise of water in capillary tubes, for example,

22
and is explained by the attractive forces between the glass walls of the tube and the
water. The glass attracts the adjacent water molecules more strongly than do other
water molecules themselves. The water is therefore "pulled" up the inside of the tube
(Figure 4.9). This "pull" is a tension which, in the terms that we are using, is a
negative (gage) pressure. That is, the gage pressure in unsaturated soils is negative.
Negative pressure heads are developed in unsaturated rock and soil matrices. The
height of rise in a capillary tube (a measure of the negative pressure head) is inversely
related to the diameter of the tube. Water will rise higher in a tube with a small diameter
than it will in a tube with a large diameter. This observation translates to soil physics in
that smaller diameter pores retain water against higher suctions than do larger pores (cf.
the large capillary tube and the small capillary tube in Fig 4.9). Thus, when water drains
from a soil or rock, large pores empty first because it takes relatively less applied suction
to pull water out of larger pores. The negative pressure produced by capillary forces,
when divided by ρg, is referred to as the capillary-pressure head.

Figure 4.9: Surface tension "pulls" water up into capillary tubes. Water pressure within the tubes
is less than atmospheric pressure, or is negative in gage units. The height of water above the free
surface in the tube is equal to the negative of the capillary-pressure head. The amount of negative
pressure head with which a capillary tube can "hold" water is inversely related to the diameter of the
tube. That is, small-diameter tubes (and by analogy, soil pores) hold water at a more negative
pressure head than do large-diameter tubes (or soil pores).

4.5.3.3 Capillary-Pressure Head and the Moisture Characteristic

For areas where there are moderate fluxes of water through the vadose zone, the two
major driving forces on soil water are the gradients in the negative capillary-pressure
head and the gradient in elevation head. This situation is exactly analogous to that for
flow in the saturated zone, with the only change being that the positive pressure heads
encountered in groundwater are replaced by negative capillary-pressure heads. Thus,
the hydraulic head for the vadose zone is defined to be the sum of the head due to
gravity and the (negative) capillary-pressure head. (In general, several forces act to
create the negative pressure heads in the unsaturated zone. The treatment given here
remains valid, but an "equivalent" negative pressure head that incorporates all important
forces, rather than just capillarity, is used. The material presented is strictly valid for
relatively moist soils and rocks and is valid for almost all circumstances with the
extended definition of negative pressure head.)

23
A suction (or negative pressure relative to atmospheric pressure) must be applied in
order to withdraw water from the unsaturated zone above the water table. The greater
the applied suction, the more water is withdrawn, and the lower is the soil-moisture
content when the soil has reached equilibrium with the applied suction. The relationship
between the external suction applied to a rock and the amount of water per bulk volume
(the moisture content) that the rock retains against that de-watering suction is called the
moisture characteristic. The applied negative pressure is a measure of the water-
retaining forces of the soil and represents the capillary-pressure head, Ψ. The moisture
characteristic generally is presented as a plot of Ψ versus θ.
The moisture characteristic for a porous material can be determined using a pressure
plate apparatus (Figure 4.10). The rock (soil) sample sits on a
porous plate made of a fine-grained material (e.g., a ceramic) that remains saturated
even at high negative pressure heads. The sample is allowed to equilibrate at a given
negative pressure head and the moisture content associated with this capillary-pressure
head is determined. The experiment is repeated at different negative heads to obtain
other points on the moisture characteristic. The locus of all such points then defines the
moisture characteristic. The moisture characteristic is one of the
important curves that define the relationships among hydraulic variables in the soil-water
system. Another is the relationship between moisture content and hydraulic conductivity.

Figure 4.10: A pressure plate for measuring capillary-pressure head. The rock sample is
placed in contact with the ceramic plate which is saturated with water under a negative
pressure head that is set by the distance of the free water surface to the right of the
diagram below the ceramic plate. The moisture content of the sample is recorded after the
sample has come to equilibrium with the selected capillary-pressure head. This
measurement gives one point on the moisture characteristic curve.

24
Figure 4.11 Moisture characteristic for a fine sand determined by starting at saturation
and draining the sample. Note that for this sand, saturation is maintained for capillary-
pressure heads between 0 and -0.36 m. The capillary fringe in such a material would be
0.36 m high. As the capillary-pressure head is reduced from about -0.40 m to -0.45 m, the
moisture content drops sharply from the saturation value of 0.35 to about 0.15. This steep
drop is typical for sandy soils. Much of the pore space is in large pores which drain once a
critical suction is exceeded. The moisture content therefore drops abruptly. On the other
hand, moisture content drops by only about 0.08 as capillary-pressure head drops from -
0.45 m to -0.60 m.

The relationship between soil water content (moisture content) and water pressure in the
pores (capillary pressure-head) is shown for contrasting soils (Figure 4.12 (a). The finely
saturated clay soil has higher initial water content owing to a higher porosity, but with
small or moderate soil water tensions, the sandy soil will release more water from its
larger pores. The withdrawal of water from a soil due to increasing tension (the drying
process) is called desorption. The reverse wetting process, the addition of water to an
unsaturated soil is called sorption. The wetting curve does not follow the same
relationship as the drying curve as shown in Fig 4.12b. The water content – pore water
pressure relationship can be very complex and a family of curves may be obtained for
varying initial states of the soil. The effect is caused by the interaction of the soil pores,
the influence of entrapped air and the changing of soil volume due to shrinking and
swelling.
Thus in determining the state of water in the soil, the soil water content θ and the soil

water tension φp should be determined since there is no single unique relationship


between these two variables.The soil moisture characteristic exhibits hysteresis- a non-
unique relationship which depends on the previous history of wetting and drying.

25
Water content,θ
Water content, θ
Desorption

(Drying/Draining) Sorption
Clay Soil
(Wetting)
Sandy Soil

Pore water pressure, φp Pore water pressure, φp

Fig 4.12a: Soil water retention Fig 4.12b Wetting and drying curves
with increasing suction The’ Hysterisis Effect’

Note: Soil scientists sometimes express tension head in pF units, where pF=log10 (-Ψ)
and Ψ is in cm of water. (Note the analogy with pH)

The Hysteresis Effect

A major factor in the explanation of hysteresis is the so-called ‘’ink-bottle’’ effect, which
arises due to the fact that many pores have relatively narrow inlets as shown in Fig 4.13.
r1
r2

Fig 4.13 The’ ink bottle’ effect


To empty such a pore, suction has first to overcome surface tension at the narrow
γ
 2  . There will then be a discontinuity as the larger pore empties. If suction
throat  
is 
 r1 
reduced, the water surface has to be drawn up by surface tension. But to fill the pore,
the surface has to rise through the widest part of the pore. To allow this, suction must be
 2γ 
  , where r2 > r1.
reduced to a value corresponding to  r
 2 
Hence a given water content is achieved at a lower suction on re-wetting.

Air entrapment may also affect the desorption-sorption characteristic. In the sorption
phase, for complete re-wetting to occur, all air must be removed from the pore space. In
practise, some will remain trapped and this can lead to a failure to ‘’close’’ the hysteresis
loop.

26
A further contribution to hysteresis is the effect of water movement on the angle of
contact ( α ) of the meniscus. Although approaching zero for a clean glass surface, for a
rough or coated surface, α >0. When water is withdraw, α tends to decrease and when
water rises, α tends to increase. These effects reinforce hysteresis. Finally, hysteresis
may occur due to a change of state of a shrinking and/or swelling soil.

5.5.3.4 Understanding Darcy’s Law

As water is drained from a saturated soil, the large pores fill with air but the smaller
spaces remain filled with water. The water in the smaller capillaries is held by surface
tension more tightly than is the water in the large pores. In a moist but unsaturated soil,
water flows through the water-filled pores while avoiding the larger, air-filled spaces. For
a given moisture content, we might conceptualize this flow as if the air-filled pores were
filled instead with a solid (e.g., wax). We would thus have flow in an equivalent,
saturated porous medium and would expect Darcy's law to be valid, but with the
hydraulic conductivity appropriate for the equivalent medium. This is, in fact, found to be
the case; Darcy's law is valid for unsaturated soil but each moisture content corresponds
to a different equivalent saturated medium, and hence a different value of hydraulic
conductivity. We therefore write the hydraulic conductivity as K(θ) indicating that it is a
function of moisture content.

The hydraulic conductivity decreases very rapidly as the medium becomes unsaturated
(Figure 4.14).

Figure 4.14: Variation of K with θ for a fine sand. Note that the scale for K is logarithmic.
The hydraulic conductivity decreases by orders of magnitude as capillary-pressure head
drops from −0.35 m to −0.55 m (moisture content decreases from saturation to about 0.06
as seen in Figure 4.11)

Note that the hydraulic gradient driving flow in unsaturated materials is the hydraulic
head appropriate for unsaturated materials, ψ + z:

27
(4.1)

Where ψ = capillary-pressure head and z = elevation head. Note that capillary-pressure


head is a function of moisture content in this expression. Darcy's law for unsaturated
conditions can thus be written:

(4.2)

5.5.3.5 Vertical Water Movement


An appreciation of many important aspects of the movement of water in the unsaturated
zone can be gained by considering one-dimensional flow in the vertical direction. In this
case, equation (4.2) becomes:

(4.3)

or, making use of equation (4.1): :

or,

(4.4)

Equation (4.4) alone is sufficient to describe the steady flow of soil moisture. As with with
groundwater flow, description of transient or unsteady processes requires the addition of
a continuity equation to Darcy's law. For the case of vertical flow of water the
appropriate continuity equation is:

(4.5)

The left side of this equation represents the rate of change of mass in a small control
volume and the right side is the difference between the inflow rate and the outflow rate,
each expressed on a per unit volume basis. That is, the equation has the same conceptual
basis as the continuity equation that we used in earlier chapters. Combining equations
(4.4) and (4.5) gives an equation that, along with information on the relationships hips
among

28
θ, ψ, and K, describes the flow of water in unsaturated rocks. The resulting
equation, referred to as the Richards' equation, is:

(4.6)

For steady flow, the time derivative of moisture content is zero and a single integration of
the right hand side of equation (4.6) recaptures equation (4.4). That is, for steady flow the
specific discharge, q , is constant and can be calculated from equation (4.4). Of course,
equation (4.4) can be used to calculate the flux in unsteady flow as well. Under unsteady teady
conditions, qz changes with time so the calculation can be taken to give a "snapshot" of
the flow at a given time. Solutions to the Richards' equation, which accounts for time
variation explicitly, yield a complete time history of heads and fluxes for specified
conditions.

4.5.3.5 The Equilibrium Profile above a Water Table

Now consider how we would expect the moisture content profile to look in a
homogenous material above a static water table under conditions of zero vertical flux.
This would be the situation we might expect from the capillary fringe through a good part
of the intermediate zone after a prolonged period without rain. In this case, equation
(4.4) becomes:

(4.7)

Dividing through by K (we can assume that K is not zero) and rearranging, we have:

or,

(4.8)

This equation can be integrated from the water table (z = 0, ψ = 0) to some arbitrary point
in the unsaturated zone (z, ψ) so that under zero-flux conditions:

(4.9)

That is, the capillary-pressure head balances the head due to gravity so there is no
hydraulic gradient and thus (by Darcy's law) no moisture flux.

29
Equation (4.9) can be used to infer the moisture distribution if the moisture characteristic ristic
(ψ as a function of θ) is known. Clearly, the equilibrium moisture profile above a water
table has exactly the same shape as the moisture characteristic because of equation (4.9). on (
Thus, for a sandy soil with a very steep moisture characteristic at the dry end of the
curve, the field capacity represents the "nearly-constant" moisture content associated with
the steep portion of the curve (Figure 4.15). The moisture characteristic for a soil oil
having a higher clay content, however, may not exhibit such a steep characteristic and in
such cases the unambiguous definition of a field capacity is problematic.

Figure 4.13 The profile of moisture content above a water table under
conditions of zero vertical flux of water. The equilibrium profile under these
conditions has the same shape as the moisture characteristic. For a fine sand, the
moisture content drops rapidly above the capillary fringe and remains relatively
constant in the range of 0.04 to 0.08 over a sizable range of elevation. This
relatively constant moisture content is referred to as field capacity.
4.5.3.6 The Profile of Capillary-Pressure Head as an Indicator of Flow

By using the knowledge that Darcy's law implies flow in the direction of decreasing
hydraulic head, we can use measurements of capillary-pressure head to infer how water
is moving in an unsaturated soil or rock. In humid to subhumid climates, a tensiometer
can be used to measure the capillary-pressure head.

If a pair of tensiometers is installed in a soil, it is easy to see how flow direction can be
determined. The example below illustrates the procedure.

30
Example

Suppose that, at noon on a day in September, moisture content in a fine sand is


measured to be 0.25 at an elevation of 3 m above the local water table and to be 0.15 at
an elevation of 3.5 m above the water table. Assuming that the hydraulic relationships in
Figures 8.3.3.4 and 8.3.3.5 hold for the sand in question, estimate the direction of flow of
water and the magnitude of the flux.

Solution

Consider first the direction of flow.


From equation (5.3) we see that, if > 0, the flow will be downward (because the
calculated qz will be negative). Conversely, if < 0, the flow will be upward.
We recognize that this conclusion merely reiterates the main point of Darcy's law, that
water flows down a gradient in hydraulic head. So, for our case, we approximate the
derivative by a finite difference:

where the subscripts refer to conditions at the different elevations above the water table.
Evaluation of the expression requires values for ψ.
From Figure 8.4, we determine that for a moisture content of 0.25, the capillary-pressure
head is about −0.42 m and for a moisture content of 0.15 it is about −0.45 m.
The calculated hydraulic gradient is 0.94 (positive) so the water flow is downward.

To estimate the magnitude of the flux,


We multiply the gradient by the hydraulic conductivity. We know that hydraulic
conductivity varies over the 3m to 3.5m interval because moisture content varies. We
might determine an "average" value of hydraulic conductivity in different ways.
For example, we could find K(θ = 0.25) and K(θ = 0.15) and average these. Or we might
just use K(θ = 0.20). The first method gives a value for K of about 0.000023 m s−1 and
the second gives a value of about 0.00002 m s−1. Thus, the specific discharge is estimated
to be approximately 0.00002 m s−1.
In some instances, one wants to infer direction of flow using measurements from a
single tensiometer. Is this possible? From equation (4.9), when there is no flow of
moisture in the vertical direction, the capillary-pressure head decreases directly with
depth, ψ = -z. Thus, the profile of pressure head with height above the water table is a
straight line. In the field, if there is no vertical flow, a measured capillary-
pressure head should plot on the straight line representing equilibrium. Suppose a
measurement shows that capillary pressure at a spot in a soil is more negative than it
would be under no-flow conditions. In this case we can
conclude that water flow will be upward if flow is steady. If measured capillary pressure
is lower (more negative) than the equilibrium value, the gradient in hydraulic head
relative to the water table will be upward and so we expect that water flow will be upward

31
because water is being drawn up from the water table. Similarly, if measured pressures
are greater than equilibrium values (region B in Figure 5.3.3.7), we can reason that
water is flowing downward away from the point of measurement (assuming steady flow).

Figure 4.14: The vertical profile of capillary-pressure head under conditions of


ψ = -z). If a measurement of capillary-pressure head falls
no flow is a straight line (ψ
on this line, it is reasonable to infer that water flow in the vertical at that point is
negligible. If measured capillary pressure is more negative (region A) than it would
be at equilibrium (for a fixed value of z), it is reasonable to infer upward movement
of water at that point. Conversely, if measured capillary pressure is less negative
(region B) than it would be at equilibrium, downward flow is indicated. The
inferences are exactly correct under steady flow conditions.

4.5.4 Soil Moisture Characteristics Under Natural Conditions

Moisture in a soil is dependent mostly on texture, soil structure and it can be seen that
sandy soils drain faster at low tension whereas clays soil retain moisture a lot more even
at high tension such that it is not available for plants to use. The Fig 4.14 shows
moisture available to plants at different tensions. Also see figs 5.14.

32
Fig 4.15 Typical moisture characteristics curves of clay, loam and sandy soils

(Adapted from USDA, SCS, National Engineering Handbook Sec.15Ch.1)

Soil Moisture Stress

It is defined as the combination of soil moisture tension and osmotic pressure. Osmotic
pressure is the pressure with which water moves across membranes from an area of
higher salts concentration to an area of lower concentration. Osmotic pressure hinders
the use of soil moisture by plant. The higher the osmotic pressure, the harder it is for the
plants to extract the water from the soil.

The rooting characteristics such as depth to which the roots the plant roots extend as
well as the proliferation/ density of the roots determine to a large extent the amount of
water taken up from the soil by the plant in addition to other factors such as soil moisture
characteristics. The deeper the roots can go the more access to soil moisture, therefore
during favourable growing periods, roots often elongate to deeper levels in order to be
able to use some of the reserved moisture down below during drought periods. Plant
roots vary from species to species for example an annual plant must extend its roots
down into the soil to make available water for its use where as perennial plants already
have elongated roots.

Apart from rooting characteristics, other factors also limit the amount of moisture
available to the plant for example high water table, shallow soils, impermeable formation
near the surface, fertility and soil salinity status as well as crop management practices.

Rooting Characteristics and Moisture Use of Crops


The amount of moisture available to crops is dependent on the depth to which plant
roots reach their depth of ability to spread out in terms of depth and lateral extent
sometimes refered to as their density. The best way to ensure that plants have enough
water/ moisture available to them during their life period is to alter their root system such

33
that the roots can reach as far down into the soil layers as possible. This will ensure that
even during the dry periods, the crops can still access the water in the deeper layers of
the soil. Some cultivation practices can be used to alter the root systems for example
cutting the top growth at different physiological stages as well as the cultivation and
cutting of surface roots. The soil pattern of an area is also known to affect the root
characteristics for example maize crop has been found to extend as deep as 1.5 meters
in medium to coarse soils while in fine soils roots are shallow.
Effective Root Zone is the depth from which the roots of average mature plant are
capable of reducing soil moisture to the extent that it is replaced by irrigation.

Table 4.9 Effective root zone depth of some common crops (grown on very deep,

well drained soils) Source: Gandhi, et al, 1970


Shallow rooted Moderately deep Deep rooted Very deep rooted
Rooted
Depth of root zone
60cm 90cm 120cm 180cm
Rice Wheat Maize Sugarcane
Potato Tobacco Cotton Citrus
Cauliflower Castor Sorghum Coffee
Cabbage Groundnut Pearl millet Apple
Lettuce Muskmelon Soyabean Grapevine
Onion Carrots Sugar beet Safflower
Pea Tomato Lucerne
Bean
Chilli

Soil Moisture Concept Terms (In relation to Agriculture)

Water in the soil may be classed as available or unavailable water. Available water is
defined as the water held in the soil between field capacity and wilting point.

Field Capacity: Is the water content of soil (volume fraction) after the saturated soil has
drained under gravity to equilibrium; (usually for about 2 days). It may also be defined as
the point at which the gravitational or easily drained water has drained from the soil.
Traditionally, it has been considered as 1/3 bar tension. However, field capacity for many
irrigated soils is approximately 1/10 bar tension.

When the soil is saturated, it will hold no more water. After rainfall/irrigation ceases,
saturated soil relinquishes water and becomes unsaturated until it can just hold a certain
amount against the forces of gravity. At this point, the larger pores are filled with air and
the smaller pores are filled with water and the drainage is slow. It can be determined by
ponding the soil surface and allowing it to drain for three days and then measuring the
moisture content of the soil.
Soil Moisture Deficit; is the extent to which water is drier than the Field Capacity;
difference between the amount of water in a soil when saturated and the amount of

34
water when drier. A Soil Moisture Deficit of 40 mm implies that 40 mm of water are
required to bring it to Field Capacity.
Wilting point: is the soil moisture content where most plants would experience
permanent wilting and is considered to occur at 15 bars tension.

Permanent wilting point: Is the volume of water content of the drying soil beyond which
a wilting plant will not recover when provided again with humid conditions. At this point
the water films around the soil are held so tightly that the plant roots cannot remove a
sufficient enough amount of water to prevent wilting unless more water is added to the
soil.
There is much less water available at field capacity in a sandy soil that drains quickly,
compared to loam or clay soil. The higher the clay content of a soil, the greater its
retention capability. At the permanent wilting point, a clay soil will still contain a
significant amount of water, but it is a finely divided state, held under high tension in the
minute pores and adsorbed to the surface of the clay particles. Fig 8.14 shows water at
different stages in the soil:

Readily available water: is that portion of the available water that is relatively easy for a
plant to use. It is common to consider about 50% of the available water as readily
available water. Even though all of the available water can be used by the plant, the
closer the soil is to the wilting point, the harder it is for the plant to use the water. Plant
stress and yield loss are possible after the readily available water has been depleted.

Soil water potential: Describes how tightly the water is held in the soil. Soil tension is
another term used to describe soil water potential. It is an indicator of how hard a plant
must work to get water from the soil. The drier the soil, the greater the soil water
potential and the harder it is to extract water from the soil. To convert from soil water
content to soil water potential requires information on soil water versus soil tension that
is available for many soils.

35
Oven dry

Unavailable for plants HYGROSCOPIC WATER


Ultimate
Wilting point

Permanent Available for survival


Wilting point

CAPILLARY WATER
Available for plant growth

Silt loam

Field capacity Sandy loam

Available moisture for sandy loam


Limited amount GRAVITIONAL OR
FREE WATER
available
Saturation
l
Fig 4.16a Water at Different Stages in Soil (Adapted from USDA, SCS Handbook Sec 15,1964)

Figure 4.16b: Available water in the Soil

36
4.5.5 Methods of measuring Soil water Content
4.5.5.1 Gravimetric Determination

A soil sample of known volume (Vt) is removed from the soil with a soil auger. The
sample is weighed (mt) and then dried in a special oven at temperatures between 100
and 110°C until the weight is constant (ms). The gravimetric water content is obtained
from:
mw mt − ms
θm = =
ms ms

Calculating the bulk density


ρb =ms Vt and knowing ρ w (1 g/cm3), the volumetric water
content, θ can be found from: θ =θm ρb
ρw

This method can be reliable and accurate if measurements are made precisely. However
it is destructive of the soil. A new sample must be taken for another measurement and
all subsequent samples should be similar to the original sample, thus a large enough
sampling area with homogeneous soil must be selected when siting the sampling points.
The main disadvantages are that it is time consuming and laboratory equipment are
required.

4.5.5.2 Neutron Scattering (Using Neutron Probe)

The neutron probe indicates soil moisture by detecting hydrogen in soil water. A counter
reads the number of neutrons that are reflected by the hydrogen in the soil. This number
is then used to calculate the moisture content.
A radioactive source is lowered into an augered hole in the soil; the first neutrons
emitted are impeded by the hydrogen nuclei of soil water. The collisions with the
hydrogen nuclei (in the water) cause a scatter of slowed nuclei and a detector senses
the density of these, dependent on the number of hydrogen nuclei present. A reading of
the detector of the slow neutrons gives a measure of the amount of water present in the
soil. The neutron source and detector are mounted in the probe, which is lowered down
an access tube previously set in the ground at a selected sampling point. Measurements
are made at various depths in the soil so that a water content profile is obtained. The
instrument must be calibrated for different soil types and each site should have its own
probe. The technique is ideal for catchment areas where soil moisture measurements
are required on a regular basis. Use of Neutron probes requires special licensing and
training.
4.5.5.3 The Tensiometer

Tensiometers provide a direct measure of the tenacity with which water is held by soils.
They measure the capillary potential. They can also be used to estimate the soil
moisture content. The tensiometer consists of a porous ceramic cup filled with water
which is buried in the soil at any desired depth and connected to a water filled tube to a
manometer or vacuum gauge (Fig. 4.17). Tensiometers have been called mechanical
roots since they provide an indication of how hard it is for the plant to get water from the
soil.

37
When the Tensiometer is placed in the soil where the tension measurements is to be
made, the bulk water inside the porous cup comes into hydraulic contact and tends to
equilibrate with soil water through the pores in the ceramic cup. When initially placed in
the soil, water contained in the tensiometer is generally at atmospheric pressure. Soil
water, being generally at sub-atmospheric pressure, exercises a suction which draws out
a certain amount of water from the rigid and airtight tensiometer, thus causing a drop in
its hydrostatic pressure. This pressure is indicated by the manometer or vacuum gauge.
An increase in soil-water content reduces tension and lowers the reading.

The main limitation in the use of tensiometers is the fact that at suctions of about 1 bar,
air dissolved in the water comes out of solution and the water column in the tensiometer
breaks. Therefore in practice, tensiometers are only useful up to suctions of about 0.85
bars (85 centibars), which is comparatively low tension for soils with high clay content.
As an irrigation guide therefore, tensiometers are helpful for crops needing nearly
saturated soils and crops grown in sandy or light loamy soils in which there is little water
left when the pore pressures approach this limit.

Tensiometers do not directly give readings of soil water content. To obtain soil water
content, a moisture release curve (water content versus soil tension) is needed.

Opening t o f ill
wit h wat er

Glas s t ube

Mer c ur y
Soil s ur f ac e c up

Connec t ing
t ube

Por ous c er amic


c up

Fig. 4.17: Typical Tensiometers; (i) connected to mercury manometer (ii)


connected to vacuum gauge

4.5.5.4 Electrical Resistance Blocks

These are simple, inexpensive instruments used for measuring soil water potential. A
porous block of gypsum with a pair of electrodes (Fig. 4.18) embedded is buried in the
soil and a meter used to read the electrical resistance of the moisture block. The block
may be entirely gypsum or covered with a porous material such as sand, fiber glass, or
ceramic. Meters are portable and are intended for use in reading a large number of
blocks throughout one or more fields. Since the blocks are porous, water moves in and
out of the block in equilibrium with the soil moisture. Meter resistance readings change

38
as moisture in the block changes which, in turn, is an indication of changes in the
amount of water in the soil. The manufacturer usually provides calibration to convert
meter readings to soil tension.
Electrical resistance blocks are made commercially and are a useful indicator of
irrigation need in crop production. One of the disadvantages of this method is that the
block has a long response time and is dependent on continuous close contact with the
soil. There is also a hysteresis effect between the wetting and drying curve relationships
between electrical resistance and soil moisture tension, but calibration data are usually
given for the drying curve. Gypsum blocks are not long lasting since there are solution
effects and they soon deteriorate in wet soils.Field instruments developed for the
measurement of moisture tension can also be connected to recorders and thus make
valuable monitoring devices for continuous assessment of soil conditions., but many of
them are not accurate and are therefore more adapted to agricultural and engineering
requirements rather than to scientific research of hydrological process. Fig 4.19 shows
the use of the electrical resistance block in measuring soil moisture.

Fig 4.18. Electrical resistance block

39
Fig 4.19 Use of an electrical resistance unit for soil moisture measurement

4.5.5.5 Pressure Membrane and Plate Technique


The apparatus consists of pressure plates contained in an air-tight metallic chamber
(Fig. 4.20). The plates are saturated and soil placed on them and also saturated. The
plates are then placed in the metallic chambers, sealed tightly and a controlled pressure
is applied. After a while, water starts to drip out of soil until equilibrium against the
applied pressure is achieved. The moisture content of the soil sample is then measured
at different pressures. The moisture contents can be plotted against the pressures to
obtain soil moisture characteristic curves.

40
Fig 4.20: Pressure plate apparatus

4.5.5.6 Time domain reflectometry (TDR)

This is a technique based on sensing the dielectric constant of the soil which is
dependent on the soil moisture. The equipment consists of an electronic meter
connected to two rods placed into the ground. The instrument sends an electrical signal
through the soil and the rods serve as the transmitter and receiver.
The TDR method uses the fact that the dielectric constant of a soil is almost completely
determined by the moisture contained in the soil. (You may recall from a physics course
that the dielectric constant is just the ratio of capacitance of a material to capacitance of
a vacuum. The dielectric constant of a soil can be measured by sending an electric wave
down parallel steel rods and analyzing the timing of the wave reflected from the open
end of the rods. The details of the physics behind the measurements can be found in
Topp et al. (1980).
4.5.5.7 Feel Method

A soil probe is used to sample the soil profile. Soil moisture is evaluated by feeling the
soil. Then a chart is used to judge relative moisture levels. It is important to sample
numerous locations throughout the field as well as several depths in the soil profile. This
method is only an estimate and lacks scientific basis.

Accurate measurement is not possible, but rather the method is an art developed over
time with extensive use. Another measurement method such as tensiometers or
resistance blocks is really needed as a reference, especially during the learning period.
The feel method requires no investment other than a soil probe.

Effective use, however, does require more time and judgment than other, more
quantitative methods. Do not use the feel method as an excuse to avoid using
tensiometers or resistance blocks.

41
4.5.5.8 The Capacitance Probe

This is another newer instrument. A nonradioactive probe is lowered into an access tube
similar to that used by the neutron probe. An electronic meter senses the amount of
moisture in the soil based on its electrical properties.

4.6 References

1. E.M Shaw, Hydrology in Practice, Chapman and Hall, 1994, London, UK.
2. J.O Ayoade, Tropical Hydrology and Water Resources, Macmillan, 1998,
London,UK.
3. E. M. Wilson, Engineering Hydrology, 4th Edition, Macmillan, 1996, London, UK
4. A. M. Michael, Irrigation Theory and Practice, Vikas Publishing House PVT Ltd,
2003, New Delhi, India.
5. S.K Garg, Hydrology and Water Resources Engineering, Khanna Publishers, 1998,
Dehli, India.
6. K. Subramanya, Engineering Hydrology, 2nd Edition, Tata McGraw-Hill Publishing,
2001.
7. W.H. Green, G.A. Ampt, Studies on Soil Physics, part 1, the flow of air and water
through soils., Journal Agricultural Sciences, 1911, Vol 4 No pp 1-24.
8. R.G.Mein, C. L. Larson, Modeling Infiltration during a Steady Rain, Water
Resources Research, 1973, Vol 9, No2 pp384-394.
9. W. Viessman,Jr, G.L. Lewis, Introduction to Hydrology, Fourth Edition, Harper
Collins College Publishers, 1996, Florida USA.
10. H.M. Raghunath, Hydrology, Principles, Analysis, Design, New Age International
Publishers, 2006, India.
11. W.J. Rawls, D.L. Brakensheik, N. Miller, Green- Ampt Infiltration parameters from
soils data, Journal Hydraulic Division, ASCE,1983, Vol 109, No 1 pp 62-70.
12. R.H. Brooks, A.T. Corey, Hydraulic Properties of Porous Media, Hydrology Papers
No 3 Colorado State University, 1964,Fort Collins, Colorado , USA.
13. B.P. Ghildyal, R.P.Tripathi, Soil Physics, New Age Publishers Ltd, 2005, New Dehli,
India.
14. V.T.Chow, D.R. Maidment, L.W. Mays, Applied Hydrology, McGraw Hill, 1988
New York, USA
15. United States Department of Agriculture, National Engineering Handbook, Section
15. Ch. 1 , Soil Plant Water Relationships. USDA, SCS, 1960, Washington D.C. USA.
16. R.T. Ghandi, P.C. Gupta, A.P. Joseph, N.D. Rege, J.R. Coover, D. F. Jones, J.T
Phelan. E.J. Pope, Handbook on Irrigation Water Requirements, Water management
Division, Department of Agriculture, Ministry of Agriculture, New Dehli, India
17. Topp, G. C., J. L. Davis, and A. P. Annan. 1980. Electromagnetic determination of soil
water content: measurements in coaxial transmission lines. Water Resources Research
16:574-582.
18. Hornberger G. M., Raffensperger J. P., Wiberg P. L., and Eshleman K. N.,
1998; Elements of Physical Hydrology; The Johns Hopkins University Press, Baltimore

42
5.7 Questions

1. What is infiltration capacity?


2. What are the factors that affect the infiltration capacity of a soil?
3. Describe two methods of determining the infiltration capacity of a soil?
4. What do you understand by the terms?

i) Field Capacity ii) Soil Moisture deficit

iii) Root Constant? Iv) Permanent Wilting Point

5 a) Distinguish between
i) ϕ -index and W-index
ii) Infiltration capacity and Infiltration rate
b) The average rainfall over 45 ha of watershed for a particular storm was as
follows:

Time (hr) 0 1 2 3 4 5 6 7
Rainfall (cm) 0.00 0.50 1.00 3.25 2.50 1.50 0.50 0.00

The volume of runoff from this storm was determined as 2.25 ha-m. Establish the ϕ -
index
6. An infiltration experiment was carried out on a small plot having sandy soil. The
resulting infiltration data was given in Table 7.9 as:
Table 7.9
Time (h)
1 2 3 4 5 6 7 9 11 13 15 17 19 21
Infiltration Capacity,
f (cm/h) 3.3 2.9 2.5 2.1 1.9 1.6 1.5 1.42 1.31 1.24 1.23 1.2 1.18 1.18

a) Derive and plot the infiltration curve


b) Derive the equation for the curve in a) above

7. Distinguish between the sorption and desorption process in the relationship of


soil water content and pore water pressure of a soil. Explain the difference
between a sandy and clayey soil.

8. Define soil moisture tension and soil moisture stress. Indicate the relationship
between them and how they affect plant growth

9. Describe briefly one method of measuring:

i) Soil moisture content


ii) Soil water Potential

10. Describe the Pressure membrane and plate technique for determining soil moisture
potential.

43

You might also like