Vous êtes sur la page 1sur 29

Johnson, C. E., and A. C. Bailey. 2002. Soil compaction.

In Advances in Soil Dynamics Volume 2,


155-178. St. Joseph, Mich.: ASAE.

Chapter 3
Soil Compaction
C. E. Johnson

A. C. Bailey

Soil is a material that is common to everyone in so many ways. We see it daily, walk
on it, travel over it, build structures on it, and use it as a medium to grow food and fiber.
So soil serves many purposes in our daily lives. Yet with all its commonness, utility and
seeming simplicity, it is a complex multiphase system with solids, liquid, and gas. At
times, soil may behave like a solid, at other times be more like a liquid, and when it
becomes airborne as dust in the wind, it flows with the air.
Throughout the world, we depend on the food and fiber produced from soil for life and
commerce. We traffic over the soil either by ourselves, or with the aid of animals or
machines, to perform the cultural practices needed to produce food and fiber and
accomplish other tasks. Sometimes, while using the soil, we abuse it and may reduce its
productivity on a short or long term basis. For example, adverse compaction may occur
when soil is trafficked or tilled at high moisture content. According to Soane and van
Ouwerkerk (1994), soil compaction is responsible for the degradation of an estimated 83
million ha worldwide. For sustainable productivity, we must manage soil compaction
during crop production in ways that the soils are not permanently harmed. Agricultural
soils are subjected to surface loads during each crop production cycle. The nature of these
loads and the specific reaction of a soil to the loads are of particular interest when one
tries to manage soil compaction.
Lyasko (1982) found that soil compaction by vehicles reduced the yield of a number
of crops including barley, wheat, oats, potatoes, and peas. He suggested a threshold index
for vehicle compaction that could be used for the design and use of field machines in crop
production. Crop roots are surrounded by soil and the chemistry, biology, and physical
structure of this soil defines the environment influencing root functions and growth
(Dexter, 1987a,b; Lowery and Schuler, 1991; Atwell, 1993).
In 1971, the American Society of Agricultural Engineers (ASAE, 1971) published a
reference book on soil compaction of agricultural soils and in it Harris (1971) presented
the collective wisdom of that time on the compaction process in agricultural soils. Yet,
soil compaction continues to be a challenge to production agriculture, particularly since
some of the large field machines have axle loads in excess of 10 t per axle (Schuler and
Wood, 1992; Soane and van Ouwerkerk, 1994). Excessive soil compaction may adversely
affect crop yields, increase tillage energy requirements, accelerate erosion, and cause
inefficient use of water and nutrients due to slow subsoil drainage. Additional field
operations and energy may be required to remove unwanted soil compaction. Hkansson
and Reeder (1994) reviewed the literature on the persistence of subsoil compaction and
crop response to subsoil compaction caused by high axle loads and concluded that
management of soil compaction is of worldwide concern.

156

Soil Compaction

Some of the recent literature on the soil compaction process, soil compaction
modeling (constitutive equations), soil compaction prediction, and management of soil
compaction will be reviewed here. Where comprehensive reviews exist, they will be
referenced but not discussed in detail. The reader is referred to the referenced literature
for further detail.
What is Soil Compaction?
The skeletal structure of soil consists of solid-like material (mineral and organic
matter) and voids which can be filled by air, water, or both. A water-saturated soil would
have its voids completely filled with water, whereas an oven-dried soil is considered to
have its voids filled with air. A soils dry bulk density (mass per unit bulk volume)
increases during compaction. When dry bulk density, b, increases, the dry bulk specific
volume, b, decreases, since b = 1/ b, which indicates that the void space has been
reduced. Gill and Vanden Berg (1967) define soil compaction as:
Soil compactness is a static state property of soil. Soil compaction changes
the state of compactness. For a specific soil, the material properties generally
do not change when the state of compactness is changed; only the static state
changes. Since soil material and state properties determine behavior
properties, a change in state of compactness indicates a probable change in
behavior properties. Hence, regardless of intended use, the soil is affected by
compaction.
The principal agents causing soil compaction are forces applied to the soil. These
forces may be created either by humans or nature. Thus, the process of compaction of soil
may be viewed as the soils behavioral reaction to compressive forces applied by nature
or humans. This fundamental soil behavior needs to be understood so it may be
characterized, quantified, and managed for the betterment of humankind.
Implications of Soil Compaction
As previously stated, an agricultural soil typically consists of a mixture of mineral
solids, water, air, and organic matter (Hillel, 1980a). The mineral solids are composed of
particles of varying size, commonly classified as sand, silt, and clay. A bulk volume of
soil, Vt, may consist of a volume of solids (minerals and organic matter), Vs, a volume of
water, Vw, and a volume of air, Va, with Vt = Vs + Vw + Va. The volume of water plus the
volume of air is the volume of voids, Vv; Vv = Vw + Va. When the soil is saturated Va is
assumed to be zero and the void volume, Vv, is filled with water. The porosity, , is an
index of the relative pore volume in the soil defined as the ratio of Vv to Vt, which
generally ranges from 0.3 to 0.6 (Hillel, 1980a). The void ratio, e, also an index of pore
volume, is defined as the ratio of the void volume, Vv, to the volume of solids, Vs. The
void ratio generally varies between 0.3 and 2.0. The void ratio is more commonly used in
engineering since it has the advantage of a denominator that is a constant.
The void volume is the sum of the volumes of all the soil pores. Many of the soil pores
create a network of micro pipes connected together throughout the soil that permits
movement of both air and water into, out of and within the soil. The quality of this
network of pore space and its structural integrity is often referred to as soil structure.
When soil is compacted, the volume that typically decreases the most is the void
volume, Vv, which, because of the soils plastic nature, does not rebound much after
mechanical forces are removed. Also, compaction of a soil alters and may damage soil
structure, the network of interconnected pore space. A compacted soil has less volume to
store water than the same soil when less compacted. The reduction of void volume and

Advances in Soil Dynamics, Vol. 2

157

alteration of soil structure reduce the soils porosity, infiltration capacity, and
permeability, which in turn affect the flow of water and air solutes into and throughout
the soil.
In general, plant growth is dependent on rooting ability, nutrient status and
accessibility of roots to nutrients, soil aeration, and water availability. Soil aeration and
available water may be subdivided into two regimes, the inter- and the intra-aggregate
regimes, which depend on the pore continuity as well as the hydraulic and concentration
gradient properties of the soil (Horn, 1990). A compacted soil has less water available to
sustain plant growth. The plant may have to expend more energy to extract the water
from the soil due to decreased pore size or lack of pore continuity caused by compaction.
Decreased pore size tends to increase the capillary forces retaining the soil water.
Consequently, crop yield may suffer if the soils void space is not frequently recharged
with water naturally (by precipitation) and/or artificially (by irrigation). If there is plenty
of soil water available for plant growth and aeration of the soil-plant root system is
sufficient, crop production may not suffer from soil compaction. However, soil
compaction may impede proliferation of the plant root system due to excessive
mechanical impedance (Atwell, 1993), thereby further decreasing the effective volume of
soil from which a plant can extract water and nutrients and reducing the plants drought
resistance. Some effects of heavy axle loads on soil air porosity, bulk density and air
permeability were investigated by Wood et al. (1993).
During mechanized crop production, the soil is trafficked by field machinery. Traction
must be developed by the prime movers, typically tractors, which pull and operate the
various field machines. Traction and mobility typically improve as a soil is compacted,
which is why roadbeds are compacted. As a soil becomes compacted, its ability to resist
additional compactive forces increases. A compacted soil requires more energy to till
than the same soil when less compacted. Soil conditions that promote plant growth and
production often conflict with the conditions necessary for efficient traction and mobility.
These contrasting needs of plant and machine must be considered in the design,
development, and management of machinery systems for crop production.

Identification and Quantification of Soil Compaction


When a soil experiences a compressive force system its specific volume typically
decreases (compressive volumetric deformation). When the force system is removed, the
amount the specific volume recovers (expansive or tensile volumetric deformation)
towards its initial condition may depend on its past history or condition prior to
application of the compressive force system and the length of time the force system was
applied. Thus, the soil can exhibit both recoverable volumetric deformation (elastic) and
permanent volumetric deformation (plastic). The force system experienced by the soil
may create shear type deformation so the volume of soil permanently changes shape
(distorts) either with or without visible formation of shear planes. If the volume changes
shape (distortion) without a volume change then the soil is said to experience plastic
flow. If the change in shape is accompanied by an increase in volume (expansion) then
the soil has typically experienced shear failure (Gill and Vanden Berg, 1967).
Measures of Soil Compaction
Some common direct measures of a soils state of compactness include dry bulk
density, b, dry bulk specific volume, b, void ratio, e, and porosity, . Indirect measures
often rely on an increase in strength or a reduction in pore space when soil compactness

158

Soil Compaction

increases. Some indirect measurements are cone penetration resistance, which reflects a
soils resistance to penetration, and permeability to air or water, which reflects the pore
space and the interconnectivity of pores. Some non-intrusive indirect methods of
measuring soil compaction, such as ground penetrating radar, neutron scattering and
transmissibility (e.g., X-ray and CT-scan) techniques often depend on both soil density
and moisture (Freeland et al., 1996).
When using indirect measures, one must be aware that other changes in the soils
condition may be incorrectly interpreted as an increase in bulk density. For example,
increased cone penetration resistance may come from changes in moisture content and
soil structural changes in aggregate bonding while the total void space or soil
compactness has not changed. Also, a soils permeability to air can change due to
moisture content or plastic flow, which disrupts the connected pore space, without any
change in dry bulk density of the soil. Indirect measures tend to assess changes in a
behavioral response highly correlated with soil compactness, such as penetration
resistance or permeability, whereas direct measures tend to assess the state of soil
compactness. The intended use of the measurement of soil compaction may dictate which
measure is most appropriate: a behavioral property or a state property.
Regardless of the chosen measure of soil compaction, three-dimensional spatial
variation of soil compaction is commonly observed in the field. Spatial variation of soil
compaction indicates either spatial variation of the compressive forces causing
compaction or spatial variation of the soils composition, structure, and prior condition,
or both. Multiple measurements within the spatial expanse of the soil are often required
to adequately characterize soil compaction. Also, the spatial variation of the soils
composition, its prior condition, and the compressive forces are often required to relate
cause and effect.
Soil Compactibility
The range of dry bulk density or void ratio that a given soil may experience is often of
interest. Both experimental and theoretical methods are available to quantify the extreme
values of bulk density. One experimental method of determining the maximum density of
a soil is an American Society for Testing and Materials standardized test, the Proctor
Density Test, described in detail in many soil mechanics textbooks and laboratory
manuals. The Proctor Density Test provides the maximum dry bulk density that can be
achieved by a given energy input at a particular moisture content.
Gupta and Larson (1979) developed a computer model to predict soil bulk density
based on particle arrangement geometry, particle density, and organic matter content.
Some basic soil input data required include particle size distribution, particle density, and
organic matter content. Spherical particles were assumed in the packing model. Output
from the model includes estimates of the minimum and maximum dry bulk densities
based on packing geometry for a particular particle-size distribution. This technique can
be useful in assessing the potential soil compatibility of a mixture of two or more soils
with different particle size distributions (such as different soil horizons, i.e., topsoil and
subsoil) (Johnson et al., 1983).

Advances in Soil Dynamics, Vol. 2

159

Basic Elements and Parameters


of Soil Compaction Behavior
Soil behavioral models capable of accurately predicting the effects of force systems
applied, either naturally or manually, would be tremendously beneficial in our quest to
preserve or conserve our soils productivity by scientifically based management for future
generations. For example, accurate prediction of effects of trafficking the soil in a
particular state or condition would be a powerful tool for production agriculture. Such
models would be valuable to engineers as they face the continuing challenge of designing
and developing effective and efficient agricultural machines. Also, models could allow
farmers to make management decisions based on the tradeoff between timeliness and
future cost of reclaiming damaged soil structure.
Development of behavioral models requires description of the inputs and outputs.
Soil compaction outputs have been traditionally defined in relation to volume change:
dry bulk density, dry bulk specific volume, porosity, void ratio, and volumetric strain
(either engineering volumetric strain, v, or natural volumetric strain, v ) (Gill and
Vanden Berg, 1967). Use of volumetric strain as an output inherently suggests that soil
is assumed to be a continuum. Vanden Berg et al. (1958) discuss the implications of
assuming soil a continuum, and Koolen (1994) has provided a review of strain theory as
applied to the mechanics of soil compaction. If the assumption of a continuum is
accepted, then the internal forces in the soil, caused by some external force system such
as a wheel on the soil surface, can be described as a continuous distribution of forces over
an internal area of soil. The validity of considering soil as a continuum is somewhat a
function of the scale of observation. From a distance, a porous soil appears to be a solid;
but up close the void space, soil aggregates, and particles may be distinguishable.
Assuming soil to be a continuum permits the use of limits, a mathematical concept and
tool necessary for the definition of the concept of stress. The concepts of both stress and
strain are presented in most introductory engineering textbooks on the mechanics of
materials, and will be reviewed here briefly. Internal stress is defined relative to an
incremental area, Ai, on a plane passing through the continuum and the normal force, Fin,
and tangential force, Fit, components acting on that area so that:
i = lim

Fin
Ai

(3.1)

i = lim

Fit
Ai

(3.2)

Ai >0

A i >0

where i represents a normal stress, and i represents a shear stress. Thus, normal and
shear stresses exist at a point (Fletcher, 1985). If the location or orientation of the cutting
plane passing through the continuum changes, then and may change in magnitude and
direction, so both and are functions of both location and orientation of the cutting
plane through the material. Figure 3.1 presents a diagram of the stress state at a point in a
continuum. Stress at a point belongs to a group of physical quantities known
mathematically as second-rank tensors and may be represented as 3 3 matrices.
Continuity considerations require that only six of the nine components are independent so
stress at a point is a symmetric matrix or tensor (Malvern, 1969) with six independent

160

Soil Compaction

components. These six components of the stress tensor are generally expressed as x, y,
z (the normal stresses), xy, xz,, and zx (the shear stresses), as illustrated in figure 3.1
and equation 3.3:
xx

xy

xz

xy
yy
yz

xz

yz

zz

(3.3)

Figure 3.1Stress state at a point in a continuum.

A characteristic of a 3 3 symmetric matrix is that its coordinate system may be


rotated to find a set of orthogonal planes on which there are no shear stresses. The planes
with no shear stresses are known as principal planes and only three non-zero components
of stress, as shown in equation 3.4, remain and are needed to describe the state of stress.
These diagonal components are known as the principal stresses.
1

0
0

0
2
0

0
3

(3.4)

Associated with any stress matrix there are three eigenvalues, the magnitude of the
principal stresses, and three eigenvectors, the orientations of the planes on which the
eigenvalues (principal stresses) act. Thus, the stress state can be expressed in terms of the
three principal stresses, 1, 2, and 3, and their directions. Figure 3.2 presents a diagram
of the principal stresses at a point.

Advances in Soil Dynamics, Vol. 2

161

Other tensor properties of stress include three invariant quantities, referred to as the
first, second, and third invariants of the stress tensor. In terms of the principal stresses,
these are:
I 1 = 1 + 2 + 3

(3.5)

I 2 = 1 2 + 2 3 + 3 1

(3.6)

I 3 = 1 2 3

(3.7)

and

Figure 3.2Principal stresses at a point.

Figure 3.3Octahedral stresses on octahedral plane.

162

Soil Compaction

A plane with its normal making equal angles to the three principal axes is known as an
octahedral plane. The normal and shear stresses that must act on this plane for
equilibrium are the octahedral normal stress, oct, and the octahedral shear stress, oct (fig.
3.3). Mathematically, they are defined, in terms of the principal stresses, as:
oct = (1 + 2 + 3 ) 3

(3.8)

2
2
2
oct = (1 2 ) + ( 2 3 ) + (3 1 ) / 3

(3.9)

and

The octahedral normal stress is also called mean normal stress. The three principal and
two octahedral stresses and their directions are independent (invariant) of the orientation
of the original three mutually orthogonal planes (x, y, and z directions) in the material.
The links between applied stresses and soil behavior, such as the compaction process,
are constitutive equations. Stress-strain relationships are the constitutive equations
important for prediction of soil compaction and yield. In their classic position paper,
Vanden Berg et al. (1958) presented the theoretical background for their hypothesis that
soil compaction was probably governed by the mean (octahedral) normal stress.
However, Shne (1958) considered the major principal stress, 1, as the dominant stress
controlling compaction.
In 1958, Roscoe et al. proposed a conceptual three-dimensional yield diagram to
describe the behavior of a saturated clay. This diagram (fig. 3.4) represented the

Figure 3.4Conceptual diagram of critical-state surface showing normal consolidation line (NCL) and
critical-state line (CSL).

Advances in Soil Dynamics, Vol. 2

163

interaction between applied effective stresses and the specific volume, v, of a saturated
clay. The two independent stress variables used were:
p = (1 +23 ) / 3
q = (1 3 )

(3.10)

where 1 = major principal stress and 3 = minor principal stress.


Roscoes data were from triaxial tests where the intermediate principal stress, 2, was
equal to the minor principal stress, 3, called a cylindrical stress state. The stress variable
p is just the octahedral normal stress for this special stress state. The stress variable q is
twice the maximum shear stress:
max = (1 3 ) / 2 = q / 2 = 3 oct / 2 2

(3.11)

Roscoes dependant variable specific volume was defined as the volume of soil contained
in a unit volume of solid material, so that v = 1 + e = (Vs + Vv)/Vs. This is different from
the definition of dry bulk specific volume presented earlier but is common in civil
engineering literature. These concepts were further developed into a unified theory called
critical state (Schofield and Wroth, 1968) as reviewed by Wulfsohn (1994) and in
Chapter 1 of this volume (pp. 48-62).
The normal consolidation line (NCL) in figure 3.4, representing specific volume as a
function of effective normal stress with no shear stress, was assumed by Roscoe et al.
(1958) to be logarithmic:
v = - ln p
(3.12)
Also, the specific volume, vk, for the elastic rebound curve (during unloading and
reloading) was assumed to be logarithmic:
v k = k - ln p

(3.13)

where and are the slopes of the respective specific volumes versus ln p and with
and k as the intercepts (at p equal to 1 unit of pressure, ln p = 0), respectively. Others
have used similar logarithmic relationships for the void ratio or dry bulk density, in place
of the specific volume in equations 3.12 and 3.13, for unsaturated soils (Larson et al.,
1980; Gupta and Larson, 1982; Desai and Siriwardane, 1984).
An important component of critical-state theory is the critical-state line (CSL in fig.
3.4), which was hypothesized to project onto the (p, q) plane as a straight line with slope
M and zero intercept (eq. 3.14), as:
(3.14)
q=Mp
A critical-state yield function defines a stable state boundary surface (SSBS) in threedimensional space (p, v, q). Britto and Gunn (1987) defined the SSBS for the Cam clay
model (the name Roscoe gave to his theoretical model for clay soil), in terms of the
coefficients in equations 3.12, 3.13 and 3.14, as:

164

Soil Compaction

tat
cal-S
Criti

p,

q, kPa

e
e lin

NC

kP
a

1+

Figure 3.5Cam clay critical-state surface (eq. 3.15). NCL is the normal consolidation line.

Mp
q=
( - v 0 - ln p)
-

(3.15)

which is illustrated in figure 3.5, where 0 is the specific volume for a hydrostatic
pressure p0 along the NCL.
A popular simplification is to assume that the critical-state yield function is an
elliptical surface, called the modified Cam clay model, mathematically defined as:
q = M p p0 - p2

(3.16)

where p0 is the hydrostatic pressure along the NCL to create a given specific volume, 0,
in equation 3.12. The projection of these two functions (eqs. 3.15 and 3.16) at a constant
specific volume 0 on the (p, q) plane are presented in figure 3.6.
From a mathematical, dimensional, and physical standpoint, equations 3.12 and 3.13
may be better represented as:
p
v = vi - ln
(3.17)
pi
where the subscript i represents some reference or initial condition when pi is not zero.
Therefore, and v have the same dimensions and units.

Advances in Soil Dynamics, Vol. 2

165

Cam clay
Modified Cam clay
M

q, kPa

p, kPa
Figure 3.6Plastic flow (eq. 3.14) and cap curves from Cam clay (eq. 3.15) and modified Cam clay (eq.
3.16) models projected on q-p plane at constant volume.

Chi et al. (1993d) investigated a similar relationship, the modified Cam clay model,
proposed by Wroth and Houlsby (1980) as follows:
v
p
ln = ln i
p
vi

(3.18)

Chi et al. (1993a) built on the concepts of a critical-state and cap model to develop an
elasto-plastic constitutive model based on two yield surfaces, the conical yield surface
from the Drucker-Prager yield function and a cap yield surface caused by hydrostatic
compression. Their model provided a single relationship; thus, it should be more easily
implemented in finite element analyses than other critical-state and cap models.
Bailey and Vanden Berg (1968) modified the critical-state yield diagram, as proposed
by Roscoe, for unsaturated agricultural soils. In saturated soils the specific volume and
the moisture content are directly related because air, the third component of soil, is
absent. In unsaturated soil, air is present, so moisture content and specific volume are not
uniquely related. The yield diagram presented by Bailey and Vanden Berg illustrated the
interactions among applied stresses (mean normal and maximum shear) and specific
volume (volume per unit dry mass) at a specific moisture content.
Later, Bailey et al. (1986) proposed equation 3.19 to represent the normal
consolidation line for unsaturated soil subjected to a hydrostatic stress:
_

v = (A + B h )(1 e C h )
_

(3.19)

where v is the natural volumetric strain; h is the hydrostatic pressure, p; and A, B, and
C are compactibility coefficients which are unique for a given soil and moisture content.

166

Soil Compaction

The hydrostatic stress state is a special stress state in which all principal stresses, 1, 2,
and 3, are equal and there are no deviatoric or shear stresses. The hydrostatic stress is
the same as the octahedral normal stress, but the term hydrostatic stress was used to
emphasize that no shear stresses were present.
Equation 3.19 represents the natural volumetric strain of an initially loose, cylindrical
soil sample under hydrostatic compression.
The mathematical form of equation 3.19 has
_
desirable qualities of simplicity, v is zero at zero h, and the stress-strain relationship
becomes linearly elastic in the limit as h approaches infinity. Figure 3.7 shows a
comparison of equations 3.12, 3.13, and 3.19.
Bulk density can be obtained from natural volumetric strain by using the relationship:
_

v = ln (V/Vi) = ln (i/)

(3.20)

where V, Vi, , and i are current volume, initial volume, current


_
_ bulk density and initial
bulk density, respectively. Note that by definition of v and v , v could replace the left
side of equation 3.18. Thus, the bulk density can be obtained from equation 3.20 using:
=

(3.21)
_
e v
Bailey and Johnson (1989) extended equation 3.19 to represent the cylindrical stress
state of the triaxial test as:
_
v = (A + Boct )(1 e Coct ) + D(oct oct )
(3.22)
where oct is the octahedral normal stress; oct is the octahedral shear stress; and A, B, C,
and D are compactibility coefficients. The coefficients A, B, and C have the same values
in both equation 3.19 and equation 3.22.

C ritical-State
NSDL - AU
d l

oct, kPa
Figure 3.7Comparing critical-state (eqs. 3.12 and 3.13) and NSDL-AU model (eq. 3.19).

Advances in Soil Dynamics, Vol. 2

167

The NSDL-AU soil behavior model (eq. 3.22) was developed to represent the
behavior of compactible agricultural soil when unsaturated and subjected to both
confining and shear stresses. Equation 3.22 represents the natural volumetric strain of an
initially loose cylindrical soil sample under a monotonically increasing cylindrical stress
state. It does not represent soil reaction to decreasing stress levels. It was developed from
modified triaxial tests on several soils using various stress loading paths, but contains the
restriction common to triaxial tests that the intermediate and minor principal stresses, 2
and 3, are equal. An upper boundary of the octahedral shear stress is reached when
maximum density (or minimum volume) is attained. At this point, when maximum
density is attained, the soil undergoes strain at constant volume, a condition that is
considered to be plastic flow and on the critical-state line for the soil. Any further
increase in shear stress will cause no further increase in density. Both Petersen (1993) and
Chi et al. (1993d) have reported good agreement between equation 3.22 and data from
soils other than those used to develop equation 3.22.
However, Bailey and Johnson (1996) observed plastic flow within equation 3.22 and
they suggested rearranging the equation to express oct as a function of oct :
_

oct = oct [ v - ( + oct)(1 - e Coct )] / D

(3.23)

At a given level of v in equation 3.23, oct is a maximum when oct /oct is zero.
Thus, the following equation was derived from equations 3.22 and 3.23 to give the
critical-state line inherent in equation 3.22:
-

octy = 2oct [ (1- e Coct) + ( + oct)(Ce -Coct )] / D

(3.24)

where octy is the value of shear stress at maximum volumetric strain. Equation 3.24 is the
relationship between octahedral shear and normal stresses at the plastic flow failure
criterion. It is analogous to equation 3.14 but is nonlinear.
Coefficients A, B, C, and D are determined from triaxial tests in which a soil sample is
subjected to a cylindrical stress state, when 2 = 3. Coefficients A, B, and C are
determined using non-linear curve fitting techniques using the hydrostatic portion of
triaxial data (no shear or deviatoric stress). Coefficient D is determined from the shear
stress loading part of triaxial tests. It is most easily determined from tests in which shear
loading occurs at constant octahedral normal stress (Bailey and Johnson, 1989), but can
be determined from conventional triaxial tests. Coefficients A, B, C, and D have been
determined for several soils. Figure 3.8 illustrates the SSBS in terms of void ratio as
defined by equations 3.22 and 3.24. Figure 3.9 illustrates a comparison of equations 3.22
and 3.24 with corresponding Cam clay critical-state models equations 3.12, 3.14, 3.15
and 3.16.
Pearman et al. (1995, 1996) and Pearman (1996) conducted research on the influence
of tractor wheel traffic on soil rut formation, stress state, dry bulk density, and penetration
resistance in the tire track. They measured six normal pressures within the soil with stress
state transducers (Harris, 1960; Nichols et al., 1987; Horn et al. 1992), which were placed
in the soil beneath the centerline of a tire path (figs. 3.10 and 3.11). Principal stresses and
octahedral stresses were calculated from the six measured normal pressures as illustrated
in figure 3.12.

168

Soil Compaction

ti
Cri
cal
i
te l

oct, kPa

a
-St
ne

NC
L

ct ,

kP
a

Figure 3.8Void ratio representation of NSDL-AU model. NCL is the normal consolidation line.

NSDL-AU
Cam clay

oct, kPa

Modified Cam clay

oct

oc
K

oct, kPa
Figure 3.9Plastic flow and cap curves from NSDL-AU, Cam clay and modified Cam clay models for
Norfolk sandy loam at a constant volume.

Advances in Soil Dynamics, Vol. 2

169

Figure 3.10Top view of a stress state transducer as placed in the soil (Pearman et al., 1996).

Figure 3.11Measured pressures in soil determined from a stress state transducer (Pearman et al.,
1996).

170

Soil Compaction

Figure 3.12Calculated stresses from data shown in figure 3.11. Pressure in the vertical direction is pz
(Pearman et al., 1996).

Pearman et al. (1995) successfully used equation 3.22, values of coefficients


developed from triaxial data for cylindrical soil samples in the laboratory, and peak
octahedral stresses from their stress state transducer data to predict soil density. The mean
predicted dry bulk densities from equation 3.22 were within +2% of the mean measured
dry bulk densities for both soils used in the investigation.
The coefficients , , k, and in equations 3.12 and 3.13 can be determined from
data generated by equation 3.22 as presented in figure 3.7. Data generated from equation
3.22 for oct = 0 and oct > 20 kPa can be fit with a semi-logarithmic curve with slope, ,
and intercept, . The slope, , and intercept, k, can be estimated by using data generated
from the NSDL-AU model for stresses less than 10 kPa.
The set of compaction behavior coefficients, either , , , k, and , or A, B, C, and
D, are unique for each soil and may change with changes in moisture content of an
unsaturated soil. Also, the soil must be uniformly prepared for testing because of the
small sample size. Despite these limitations, these constitutive relationships provide an
accurate determination of soil density at a given stress state in the soil profile.
A soil compaction model easily implemented in the finite element method (FEM) is
the hyperbolic model first proposed by Kondner (1963) and modified for use in FEM by
Duncan and Chang (1970). This model provides a nonlinear elastic approach to
describing the compaction process. The tangent modulus of elasticity, Et, is represented
by:

Advances in Soil Dynamics, Vol. 2

E t = K i Pa 3
Pa

171

ni

R f (1 - sin )(1 3 )
1
2 c cos + 2 sin
3

(3.25)

where 1 = major principal stress, 3 = minor principal stress, Pa = atmospheric pressure,


= soil internal friction angle, c = soil cohesion, Rf = failure ratio defined as the ratio of
the maximum failure deviatoric stress obtained from triaxial tests to the ultimate
deviatoric stress obtained from regression analysis, and Ki, ni = dimensionless parameters
obtained from tests.
Additional details on this hyperbolic equation (eq. 3.25) can be found in Volume 1 of
these Advances (p. 169), Upadhyaya (1994), or Chi et al. (1993d). Shear stress plays a
role in soil compaction, whereas classical continuum mechanics generally assumes that
volume is not affected by shear stress. Note that an increase in oct in equation 3.22
indicates increased compaction until plastic flow begins (limit of octy in eq. 3.24) and an
increase in the deviatoric stress, 1 3 or q, indicates an increase in the tangential
modulus of elasticity in equation 3.25. Shear stress also causes distortion (shear strain),
which has been modeled by Johnson and Bailey (1990), Chancellor (1993), Chancellor
and Upadhyaya (1994), and Petersen (1994). Shen and Kushwaha (1995) developed a
stress-strain-time model for unsaturated soil under shear loading conditions that may be
applicable to soil behavior near or at the soil-tire or track interface when travel reduction
occurs.
The role of the intermediate principal stress, 2, when 2 and 3 are not equal, is not
well understood and has received little attention. Dunlap and Weber (1971), Kumar and
Weber (1974), and Khan and Hoag (1978), all at the University of Illinois, have
conducted the most extensive research on the influence of 2 on the stress-strain behavior
of unsaturated soil. Gibas et al. (1993) developed a true three-dimensional triaxial test
device capable of applying all three principal stresses (1, 2, and 3) independently to a
cubical soil sample.
Also, the influence of unloading and reloading or repeated loading has received little
attention. Studies of rut formation under multiple passes of tires in tilled soil indicate that
rut depth increases with each pass at a decreasing rate with approximately 90% of the
total rut depth typically caused by the first pass (Taylor et al., 1982). Johnson et al. (1992)
studied repeated loading by various stress paths on soil using a computer-controlled
triaxial apparatus (cylindrical stress state). They found only a slight increase in bulk
density as the number of reloadings increases, if the stress path and the maximum stress
state attained during repeated loading remained the same. Koolen and Kuipers (1983)
report that if an unsaturated soil experiences kneading action (different stress paths
during reloading), an increase in bulk density will occur compared to no kneading action.
These results suggest that:
1. Multiple passes of a tire may not reload the soil along the same stress paths, or
2. The soil experiences increased maximum stress states with multiple passes of the
same tire, because the dynamic load on the tire is distributed over a smaller area
with each pass, or
3. A combination of 1 and 2.
Thus, further research is needed to determine effects of repeated loading on soil
compaction.

172

Soil Compaction

Much of the past literature on soil compaction modeling has not addressed strain or
stress rate effects. However, Adam and Erbach (1995) observed that the depth of ruts
formed by tractor tire traffic decreased and soil strains measured beneath the tire track
decreased (Erbach et al., 1991) as forward velocity increased. Johnson et al. (1972) found
that agricultural soils do exhibit load relaxation following sudden step enforced
displacements. Yong and Warkentin (1966, 1975) present basic soil behavior concepts
and some stress-strain rheological models in their soil mechanics texts. Chung and Lee
(1975) modeled the soil as a viscoplastic material in a finite element analysis of soil
behavior under a moving wheel. Oida and Yoshimura (1980) and Hiroma and Ota (1990)
assumed soil to act as a viscoelastic material, a three element Voigt model, in a finite
element analysis of soil under a wheel. Srivastava and Steffe (1987) used rheological
models to study soil compaction.
Unfortunately, none of the constitutive relationships reviewed predict structural
changes in the soil, such as changes to the network of interconnected pore space, which
have direct influence on air and water infiltration and permeability. The stress path that a
soil experiences influences the resulting shear strain or distortion (Grisso et al., 1987;
Johnson and Bailey, 1990). However, the changes to the continuity in the network of soil
pores caused by compaction, distortion, or both, are not as well understood. Blackwell et
al. (1990) studied the response of pore channels created by roots to compression by
vertical stress applied to the soil. Dawidowski and Koolen (1987) found that shear
deformation of a saturated or near-saturated soil can cause a dramatic change in hydraulic
conductivity of the soil.
Fundamental theories of soil mechanics for saturated soils have been applied to
unsaturated agricultural soils (Reece, 1977; Hettiaratchi and OCallaghan, 1980). The
soil-water tension in unsaturated soils has been shown to influence both stress-strain and
strength characteristics (Leeson and Campbell, 1983). However, integration of the
influence of moisture content on unsaturated soil behavior into constitutive relationships,
in a manner that has physical meaning (not just by empirical correlation), has been a
struggle. Hettiaratchi and OCallaghan (1980, 1985) illustrate one method. More recently,
Fredlund and Rahardjo (1993) and Wulfsohn et al. (1996, 1998) have developed a theory
for unsaturated soil mechanics behavior in which the soil is considered to have four
phases: solid, air, water, and contractile skin (the air-water interface). In an unsaturated
soil, the contractile skin acts as a rubber membrane that can induce a matric potential in
the soil pores. Future research may find a way to integrate the relationships between
moisture content, void space and geometry, and matric potential (the soil-water tension
characteristic curve developed by soil physicists and described by Hillel, 1980a,b) for a
soil into constitutive relationships of unsaturated soil behavior. Field soils may be
subjected to stresses caused by field machinery and natural forces over a wide range of
soil moisture conditions. Thus, the role and influence of moisture content on the behavior
of unsaturated agricultural soil needs to be reflected in the mathematical models of soil
compaction behavior.
Coefficients for the soil compaction models reviewed are typically determined by
regression techniques applied to data collected in the laboratory from very loose soil
samples in conventional triaxial cells (cylindrical stress-controlled testing) or in confined
compression cells (axisymmetric strain-controlled testing), and so the compaction data
represent a controlled remolding of the soil. Thus, the results may not reflect the behavior
of the parent in situ soil in the field, which may be a structured soil developed over time
with stable soil aggregates bonded together by cementing agents from chemical and

Advances in Soil Dynamics, Vol. 2

173

biological activity. Horn (1981, 1988) and Horn and Lebert (1994) are some of the few
researchers to address the compaction process in structured soils and on individual soil
aggregates. Horn used a model developed by Kezdi (1976) to predict behavior of
structured, unsaturated agricultural soil.
Further progress in development of more meaningful and comprehensive constitutive
relationships should be guided by general validity and physical realism. As Prevost
(1987) emphasized:
Further progress in expanding analytical capabilities in geomechanics now
depends upon consistent mathematical formulations of generally valid and
realistic material constitutive relations.
Useful constitutive relationships should not only help us to accurately predict soil
behavior but should also reflect an accurate description of behavioral conceptualization
(Schafer et al., 1991). Also, Schafer et al. (1992) discuss other needs in soil compaction
and soil compaction management research.
Currently, the constitutive relationships for soil behavior, as reviewed, provide a
determination of volumetric strain, specific volume, void ratio, or soil density at only a
given stress state in the soil profile. The stress state may vary throughout the soil profile
as observed from stress measurements (Harris, 1960; Nichols et al., 1987; Horn et al.,
1992). If soil behaves similarly to a solid continuum, then we would expect maximum
stresses near the load and reduced stresses away from the load. However, Shne et al.
(1962) observed that the maximum amount of soil compaction occurred somewhat below
the soil surface. Thus, it appears that some aspects of continuum theory may be too
limited for soils. Prediction of soil compaction throughout a profile requires simultaneous
prediction of the distribution of stresses throughout the profile. Continuum mechanics
concepts have been the most extensively used concepts in soil compaction prediction
techniques.
Soil Compaction Prediction
Feda (1978) describes some methods, including methods using the Boussinesq,
Cerruti, and Frlich equations, to predict the distribution of stresses throughout a soil
profile generated by contact stresses at the soil surface. The Boussinesq and Cerruti
equations assume that the soil is a homogeneous, isotropic, and elastic material. The
Boussinesq equation was developed for a concentrated point load normal to the surface.
The Cerruti equation was developed for a concentrated point load tangential to the
surface. The Frlich equation is a modification of the Boussinesq equation that assumes
the modulus of elasticity to increase with depth in the profile, dependent on the value of a
concentration factor, and a Poissons ratio of 0.5 (i.e., no volume change within the soil
mass). The Poissons ratio assumption and it implications are seldom pointed out.
Johnson and Burt (1990) used the Frlich and modified Cerruti equations to predict the
stress distribution in the soil profile beneath a powered tire from various types of pressure
distributions at the soil-tire interface with reasonable success. One could use the stresses
predicted at a point in the profile as input to one of the constitutive relationships for soil
behavior and predict a volumetric strain, void ratio, or bulk density for that point. Smith
(1984) and Van den Akker and Van Wijk (1985) used this technique to predict the bulk
density profile directly beneath tires with differing dynamic loads. However, prediction
of rut formation on the surface and soil deformation throughout the profile should be used
cautiously in view of the elastic theory and Poissons ratio assumptions and their
implications in these equations. Gupta and Raper (1994) reviewed use of this technique

174

Soil Compaction

and some axisymmetric and plane strain Finite Element Methods (FEM) for prediction of
soil compaction.
Bolling (1985) describes a method of measuring mean normal pressure under a tire
with a pressure bulb apparatus. He correlated the pressure measurements under tires with
dynamic load on the tire, tire size, and laboratory triaxial test data on a given soil to
devise a method of relating tire size, dynamic load, and soil compaction data so that tires
could be selected to minimize soil compaction potential. If one assumes soil to behave as
a continuum, then applicable numerical techniques include the Finite Element Method
(FEM) and the Boundary Element Method (BEM). FEM has been the most popular of the
two methods to date. Perumpral et al. (1971), Chung and Lee (1975), Yong and Fattah
(1976), and Yong (1984) were some of the first researchers to apply the FEM to
prediction of soil compaction under tire traffic on soil. Later, Pollock et al. (1985) used
the FEM to investigate multipass effects of vehicles on soil compaction with a customwritten finite element program.
Upadhyaya (1994) provided an excellent review of incorporating both linear and
nonlinear elastic constitutive models of soil in the FEM. Kushwaha and Shen (1994)
provided an excellent review of plasticity theory (yield and failure functions, flow rules,
and hardening laws) in soil constitutive modeling. These concepts are also reviewed in
this volume (pp. 30-48).
Both equations 3.19 and 3.22 have been used as basis for the stress-strain relationship
inputs for finite element analyses of the soil compaction process. Raper et al. (1987) used
equation 3.19 to develop a nonlinear elastic custom-written FEM model of soil
compaction. Later, Block et al. (1994) and Raper et al. (1995) predicted soil compaction
and stress distribution caused by a rigid wheel with a custom-written FEM program
developed by Raper et al. (1994). Block et al. (1994) and Raper et al. (1995) used a plane
strain FEM model assuming the wheel had infinite length along its axis. They modeled
the soil as a nonlinear elastic material with variable Poissons ratio, based on equation
3.22. Block et al. (1994) loaded the soil by incremental vertical displacement of the soil
FEM model mesh in contact with the wheel until the measured rut depth was reached.
Raper et al. (1995) loaded the soil with incremental vertical stress until the measured
maximum vertical stresses along the wheel in contact with the soil were reached.
Raper et al. (1994) modeled soil stresses and soil deformation caused by an 18.4 R38
agricultural tractor tire at various tire inflation pressures and dynamic loads. They used a
plane strain FEM model in a plane perpendicular to the direction of travel assuming the
action of the tire on the soil was like that of a foundation footing along the tire path. The
soil FEM model mesh, as illustrated in figure 3.13, was incrementally loaded up to the
measured peak soil-tire interface stresses from the tire center to the tire edge. Symmetry
was assumed about the centerline of the tire. Predicted iso-stress lines for different
inflation pressure-dynamic load combinations, like those illustrated in figure 3.14,
showed that similar soil stress patterns can develop with combinations of different
inflation pressures and different dynamic loads. Octahedral normal and shear stresses
calculated from stress state transducer data were mostly predicted within a 95%
confidence interval by the plane strain FEM model assuming the soil as a nonlinear
elastic material with variable Poissons ratio, based on equation 3.22.

Advances in Soil Dynamics, Vol. 2

175

Figure 3.13Example of original and deformed FEM mesh (Raper et al., 1994).

Figure 3.14Finite element predicted octahedral normal stress isolines for FEM mesh shown in figure
3.13 (Raper et al., 1994).

Also, Foster et al. (1995) used equation 3.19 to model soil as a nonlinear elastic
material for use in NASTRAN, a commercially-available finite element program. To
illustrate how a soil compaction behavior model can be transformed for use in a nonlinear
elastic FEM model, their development is summarized here. The engineering volumetric
strain, v, and oct in a linear elastic material are related by:
oct = E v / (3(1 - 2)) = K v

(3.26)

where E = modulus of elasticity, = Poissons ratio, and K = bulk modulus.


The tangential modulus of elasticity, Et, is generally used for nonlinear elastic
materials (a variable moduli approach) and is defined by differentiating equation 3.26
with respect to v:

176

Soil Compaction

E t = 3(1 - 2) d oct / d oct

(3.27)

Poissons ratio is assumed to be a constant. The shear stress is ignored when


developing the tangential modulus of elasticity because the volumetric strain of an elastic
material is assumed to depend on only the normal stress. From equation 3.19:
_

d v / d oct = B(1 - e-Coct ) + C(A+ B oct ) e-C oct

(3.28)

Converting from natural strain to engineering strain:


_

d v = d v/( v +1)

(3.29)

d v / d oct = ( v + 1)[B(1 - e-Coct ) + C(A+ B oct ) e-Coct ]

(3.30)

Using the definition of natural volumetric strain:


_

v = ln (V/Vi) = ln(v + 1)

(3.31)

the inverse of equation 3.30 becomes:


_

d ct / d v = e v / [B(1 - e-Coct ) + C(A+ B oct ) e-C oct ]

(3.32)

and then the tangential modulus of elasticity becomes:


E t = 3(1 - 2) e

- v

/ [B(1 - e-C oct ) + C(A+ B oct ) e-Coct ]

(3.33)

Both the soil and wheel were modeled by the FEM. The wheel was modeled as steel,
an elastic but rigid material with respect to the soil. The soil was loaded by applying an
incremental vertical force at the wheel axle until the desired dynamic load was obtained.
This approach was used with reasonable success to predict stresses and rut depth under a
rigid driven wheel reported by Block (1991). Foster et al. (1995) speculated that
prediction would be improved if a variable tangential Poissons ratio was used along with
the tangential modulus of elasticity as had been done by Raper et al. (1995) with a
custom-written finite element program.
Chi et al. (1993b, 1993c) developed one of the first three-dimensional custom-written
FEM programs to predict soil compaction. They used the hyperbolic model (eq. 3.25) in
their FEM program to predict effects of tire traffic on compaction of agricultural soil.
Later, Chi and Tessier (1994) used the modified Cam clay critical-state model (eqs. 3.16
and 3.18). Their finite element results showed significant influence of the two stressstrain models on predicted volume change, stress distribution, and disturbed soil-tire
interface profile. The critical-state model predicted the stress and volumetric strain more
concentrated directly under the tire and deeper into the soil than the hyperbolic model,
which predicted more lateral expansion of volumetric strain and stresses.
Kirby (1989) used an elasto-plastic model in a FEM program to predict damage under
tire traffic on wet clay soil due to shear. He postulated that when soil is at or near
saturation there will be very little volume change or compaction, since the water does not
have enough time to escape from the pores during passage of the tire. The soil was
assumed to fail in shear either according to Mohr-Coulomb failure theory (cohesion plus
internal friction) or a Tresca failure law (pure cohesion and no internal friction). Uniform
normal contact pressure and surface traction (shear) was applied at the soil-tire interface

Advances in Soil Dynamics, Vol. 2

177

in a plane strain FEM analysis and only the normal contact pressure at the interface was
applied in an axisymmetric FEM analysis. In both analyses a uniform soil profile was
assumed. The extent of predicted shear damage depended on the width of the tire, the
contact pressure, and the shear strength of the soil. He concluded that the axisymmetric
analysis, in which surface traction cannot be produced, underestimated the depth of
damage, whereas the plane strain analysis, in which out-of-plane shear stresses are
ignored, overestimated the depth of damage.
ABAQUS, another commercially-available FEM program, has incorporated various
elastic-plastic cap constitutive relationships to model geotechnic materials. For example,
ABAQUS permits use of the normal consolidation line for the void ratio (similar to eq.
3.12):
e = N - ln p

(3.34)

and an elastic curve for the void ratio, ek (similar to equation 3.13):
ek = Nk - ln p

(3.35)

ABAQUS also allows input of tables of values that represent components of various
elastic-plastic cap constitutive relationships. Chiroux et al. (1997) developed a threedimensional stress-strain dynamic model of a rigid wheel rolling on a soil surface using
ABAQUSs Drucker-Prager cap model with input of tabular values generated by equation
3.19 and other parameters estimated from equation 3.22 for the soils used by Block
(1991). Predicted stresses in the soil under the wheel and rut depths (without rebound)
following passage of the wheel from the three-dimensional model were in good
agreement with those observed and measured by Block (1991). However, the soil stresses
tended to be over-predicted when using the plane strain assumption. A unique feature of
Chirouxs ABAQUS FEM model is that the dynamic load on the wheel supported by the
soil is represented by an equivalent mass with downward acceleration of 1 g. Most other
FEM models for soil compaction prediction reported in the literature have been twodimensional, plane strain, or axisymmetric, and have loaded the soil by either enforced
displacement (based on measured rut depth) or an assumed or measured distribution of
stress acting on the footprint area.
For simplicity, the stresses applied over the tire footprints are often assumed uniform
when predicting soil compaction. However, research measurements have shown this
assumption to be in error (Vanden Berg and Gill, 1962). Burt et al. (1987), Wood (1988),
and Burt et al. (1992) measured maximum normal contact pressures between the tire and
the soil that were over five times the inflation pressure inside the radial tractor tire. The
location of the maximum contact pressure tended to move from the centerline of the
footprint toward the outside edge as the dynamic load increased.
The BEM has been successfully applied to inelastic materials and geomechanics
problems (Telles, 1983; Venturini, 1983). The BEM and FEM have been coupled to solve
some soil-structure interaction problems (von Estorff and Kausel, 1989). This method
may be applicable to modeling the soil-tire system since the rubber tire and soil behave
quite differently (Carstensen, 1996).
In the strictest sense, soil is not a continuum, but rather a system of soil particles or
aggregates that may be attached together by cementing agents that may or may not
behave as a continuum. Yamamoto (1972, 1975) applied particle model methods to stress
analysis in soils. An emerging soil behavior prediction technique is the Discrete Element
Method (DEM) that models the motion and stress-strain behavior of each individual soil

178

Soil Compaction

particle or aggregate. At present, the DEM definitely requires a fast computer with
lots of memory to solve any soil compaction problem of practical interest. Ting et al.
(1989) used the DEM to investigate modeling of soil behavior. The DEM seems most
useful when the material experiences shear flows or has regions of plastic flow (Hopkins
et al., 1992). There is potential to couple the DEM with the either the BEM or FEM to
model soil behavior under conditions where soil exhibits failure or plastic flow zones
within its continuum mass.
Soil Compaction Management
The management of soil compaction was a focus of the Eighth Conference of the
International Soil Tillage Research Organization (ISTRO) in 1979. Since then, soil
compaction and management of soil compaction has been a major topic at other
international conferences such as the First and Second International Conferences on Soil
Dynamics, other ISTRO conferences, and International Society for Terrain-Vehicle
Systems (ISTVS) conferences (Taylor and Gill, 1984). The distinguished lecture by
Hettiaratchi and OCallaghan (1985), at the First International Conference on Soil
Dynamics, presented the application of critical-state soil mechanics concepts to the
mechanical behavior of unsaturated soils. Another distinguished lecture by Soane (1985)
focused on management of traction and transport systems in cropping systems. He
suggested the concept of an optimum level of compaction that would maximize crop
yield. The optimum level would be dependent on the sensitivity of the crop, the soil, the
climate, and potentially the cultural practices. An international conference sponsored by
NATO focused on the mechanics and related processes in structured agricultural soils.
Gupta et al. (1989) presented a position paper describing prior research efforts on
modeling of soil compaction behavior and suggested future thrusts.
Many of the papers presented at these conferences are available in respective
proceedings. Agricultural experiment stations in some of the states of the U.S. (i.e.,
Chancellor, 1977) and other countries (i.e., Eriksson et al., 1974) have published bulletins
on causes and effects of soil compaction and management of soil compaction. Soane and
van Ouwerkerk (1994) edited a book on effects of soil compaction and management of
soil compaction in which maximum acceptable applied stress in the field is discussed.
Certain common principles for soil compaction management seem to emerge from the
literature on soil compaction behavior modeling, soil compaction prediction, and
management of compaction of unsaturated agricultural soils. These principles should
guide researchers and design engineers in modifying current production systems and in
the development and design of future crop production systems. The simplest of these
principles in management of soil compaction are:
1. The soil compaction process and the state of compactness in agricultural soils
depends on stress, rarely uniform, applied by nature, animals, humans or machines;
2. Soil compactibility depends on the soil type, its structure and moisture content, i.e.
all soils are not equal; and
3. The optimum soil compactness for a crops growth and yield and the optimum soil
compactness for energy-efficient mobility and traction usually differ.
The design concepts for construction and operation of low inflation pressure radial-ply
tires, rubber tracks, matched tread width, and controlled traffic (Taylor and Gill, 1984),
plus others, have resulted from engineers applying these principles to the design of
machinery for production agriculture.

Advances in Soil Dynamics, Vol. 2

References for Chapter 3


Adam, K.M., and D.C. Erbach. 1995. Relationship of tire sinkage depth to depth of soil
compaction. Transactions of the ASAE 38(4): 1011-1016.
ASAE. 1971. Compaction of Agricultural Soils, ed. K.K. Barnes, W.M. Carleton, H.M. Taylor, R.I.
Throckmorton and G. E. Vanden Berg. St. Joseph, Mich.: ASAE.
Atwell, B.J. 1993. Response of roots to mechanical impedance. Environmental and Experimental
Botany 33: 27-40.
Bailey, A.C., and C.E. Johnson. 1989. A soil compaction model for cylindrical stress states.
Transactions of the ASAE 32(3): 822-825.
Bailey, A.C., and C.E. Johnson. 1996. Soil critical state behavior in the NSDL-AU model. ASAE
Paper 96-1064. St. Joseph, Mich.: ASAE.
Bailey, A.C., and G.E. Vanden Berg. 1968. Yielding by compaction and shear in unsaturated soils.
Transactions of the ASAE 11(3): 307-311, 317.
Bailey, A.C., C.E. Johnson, and R.L. Schafer. 1986. A model for agricultural soil compaction. J.
Agric. Eng. Res. 33: 257-262.
Blackwell, P.S., T.W. Green, and W.K. Mason. 1990. Responses of biopore channels from roots to
compression by vertical stresses. Soil Sci. Soc. Am. J. 54: 1088-91.
Block, W.A. 1991. Analysis of soil stress under rigid wheel loading. Ph.D. dissertation. Auburn
University, Auburn, Ala.
Block, W.A., C.E. Johnson, A.C. Bailey, and R.L. Raper. 1994. Energy analysis of finite element
soil stress prediction. Transactions of the ASAE 37: 1757-1762.
Bolling, I.H. 1985. How to predict soil compaction of agricultural tires. In Proc. Intl. Conference
on Soil Dynamics 5: 936-952. Auburn, Ala.: National Soil Dynamics Laboratory.
Britto, A.M., and M.J. Gunn. 1987. Critical State Soil Mechanics Via Finite Elements. New York:
Halsted Press.
Burt, E.C., A.C. Bailey, and R.K. Wood. 1987. Effects of soil and operational parameters on soiltire interface stress vectors. J. Terramechanics 24(3): 235-246.
Burt, E.C., R.K. Wood, and A.C. Bailey. 1992. Some comparisons of average to peak soil-tire
contact pressures. Transactions of the ASAE 35(2): 401-404.
Carstensen, C. 1996. Coupling of FEM and BEM for interface problems in viscoplasticity and
plasticity with hardening. SIAM Journal on Numerical Analysis 33: 171-207.
Chancellor, W.J. 1977. Compaction of Soil by Agricultural Equipment. Bulletin No.1881,
University of California, Division of Agricultural Sciences, Richmond.
Chancellor, W.J. 1993. Generalized stress-strain characteristics for cylindrical samples during
compaction and shear. ASAE Paper No. 93-1524, St. Joseph, Mich.: ASAE.
Chancellor, W.J., and S.K. Upadhaya. 1994. Effects of stability mechanisms during triaxial tests of
cylindrical soil samples on stress ratio vs. strain ratio relations. In Extended Abstracts, Second
Intl. Conference on Soil Dynamics 18-20. Silsoe, Bedford, UK: Silsoe College.
Chi, L., and S. Tessier. 1994. Comparison of nonlinear elastic and elasto-plastic models. ASAE
Paper N0. 94-1076. St. Joseph, Mich.: ASAE.
Chi, L., R.L. Kushwaha, and J. Shen. 1993a. An elasto-plastic constitutive model for agricultural
cohesive soil. Can. Agric. Eng. 35(4): 245-251.
Chi, L., S. Tessier, and C. Lagu. 1993b. Finite element prediction of soil compaction induced by
various running gears. Transactions of the ASAE 36(3): 629-636.
Chi, L., S. Tessier, and C. Lagu. 1993c. Finite element modeling of soil compaction by liquid
manure spreaders. Transactions of the ASAE 36(3): 637-644.
Chi, L., S. Tessier, E. McKyes, and C. Lagu. 1993d. Modeling mechanical behavior of agricultural
soils. Transactions of the ASAE 36(6): 1563-1570.
Chiroux, R.C., W.A. Foster, C.E. Johnson, S.A. Shoop, and R.L. Raper. 1997. Three dimensional
finite element analysis of soil interaction with a rigid wheel. ASAE Paper No. 97-1028. St.
Joseph, Mich.: ASAE.

Advances in Soil Dynamics, Vol. 2

Chung, T.J., and J.K. Lee. 1975. Dynamics of a viscoplastic soil under a moving wheel. J.
Terramechanics 12(1): 15-31.
Dawidowski, J.B., and A.J. Koolen. 1987. Changes in soil water suction, conductivity and dry
strength during deformation of wet, undisturbed samples. Soil Tillage Res. 9: 169-180.
Desai, C.S., and H.J. Siriwardane. 1984. Constitutive laws for engineering materials with emphasis
on geologic materials. Englewood Cliffs, N.J.: Prentice-Hall.
Dexter, A.R. 1987a. Compression of soil around roots. Plant Soil 97(3): 401-406.
Dexter, A.R. 1987b. Mechanics of root growth. Plant Soil 98(3): 303-312.
Duncan, J.M., and C.Y. Chang. 1970. Nonlinear analysis of stress and strain in soils. In Proc.
American Society of Civil Engineers, J. Soil Mechan. Foundations Div. 95(SM5): 1629-1653.
Dunlap, W.H., and J.A. Weber. 1971. Compaction of an unsaturated soil under general state of
stress. Transactions of the ASAE 14(4): 601-607, 611.
Erbach, D.C., G.R. Kinney, and A.P. Wilcox. 1991. Strain gage to measure soil compaction.
Transactions of the ASAE 34(6): 2345-2348.
Eriksson, J., I. Hkansson, and B. Danfors. 1974. The effect of soil compaction on soil structure
and crop yields. Bulletin No. 544, Swedish Institute of Agricultural Engineering, Uppsala,
Sweden.
Feda, J. 1978. Stress in Subsoil and Methods of Final Settlement Calculation. New York: Elsevier.
Fletcher, D.Q. 1985. Mechanics of Materials. New York: Holt, Rinehart and Winston.
Foster, W. A., Jr., C.E. Johnson, R.L. Raper, and S.A. Shoop. 1995. Soil deformation and stress
analysis under a rolling wheel. In Proc. of the 5th North American Conference/Workshop of the
Intl. Society for Terrain-Vehicle Systems, 194-199. Saskatoon, Canada.
Freeland, R S., L G. Wells, and R.D. Dodd. 1996. Assessing soil properties for site-specific
agriculture using ground penetrating radar. SAGEEP 96 Proc. Symposium on the Application of
Geophysics to Engineering and Environmental Problems. April 28-May 2, Keystone, Colo.,
Environmental and Engineering Geophysical Society. Wheat Ridge, Colo., pp. 1115-1119.
Fredlund, D.G., and H. Rahardjo. 1993. Soil Mechanics for Unsaturated Soils. New York: Wiley
and Sons.
Gibas, D.M., R.L. Raper, A.C. Bailey, and C.E. Johnson. 1993. Cubical pneumatic cushion triaxial
soil test unit. Transactions of the ASAE 36(6): 1547-1553.
Gill, W.R., and G.E. Vanden Berg. 1967. Soil Dynamics in Tillage and Traction. Agriculture
Handbook No. 316. Agricultural Research Service, U. S. Department of Agriculture.
Grisso, R.D., C.E. Johnson, and A.C. Bailey. 1987. The influence of stress path on distortion during
soil compaction. Transactions of the ASAE 30(5): 1302-1307.
Gupta, S.C., A. Hadas, and R.L. Schafer. 1989. Modeling soil mechanical behavior during
compaction. In Mechanics and Related Processes in Structured Agricultural Soils. NATO ASI
Series E: Applied Sci. Vol. 172. Proceedings of the NATO Advanced Research Workshop on
Mechanics and Related Processes in Structured Agricultural Soils, St. Paul, Minn. London:
Kluwer Academic.
Gupta, S.C., and W.E. Larson. 1979. A model for predicting packing of soils from particle-size
distribution. Soil Sci. Soc. Am. J. 43: 758-764.
Gupta, S.C., and W.E. Larson. 1982. Modeling soil mechanical behavior during tillage. In
Predicting Tillage Effects on Soil Physical Properties and Processes, 151-178. Am. Soc. of
Agron. Special Publ. 44, ed. P. Unger et al.
Gupta, S.C., and R.L. Raper. 1994. Prediction of soil compaction under vehicles. In Soil
Compaction in Crop Production, ed. B.D. Soane and C. van Ouwerkerk, 71-90. Amsterdam:
Elsevier.
Hkansson, I., and R.C. Reeder. 1994. Subsoil compaction by vehicles with high axle load-extent,
persistence and crop response. Soil Tillage Res. 29: 277-304.
Harris, W.L. 1960. Dynamic stress transducers and the use of continuum mechanics in the study of
various soil stress-strain relationships. Ph.D. dissertation. Michigan State Univ. East Lansing.
Harris, W.L. 1971. The soil compaction process. In Compaction of Agricultural Soils, ed. K.K.
Barnes, W.M. Carleton, H.M. Taylor, R.I. Throckmorton, and G. E. Vanden Berg, 9-44. St.
Joseph, Mich.: ASAE.

Advances in Soil Dynamics, Vol. 2

Hettiaratchi, D.R.P., and J.R. OCallaghan. 1980. Mechanical behavior of agricultural soils. J.
Agric. Eng. Res. 25(3): 239-259.
Hettiaratchi, D.R.P., and J.R. OCallaghan. 1985. The mechanical behavior of unsaturated soil. In
Proc. Intl. Conference on Soil Dynamics 2: 266-281. Auburn, Ala.: National Soil Dynamics
Laboratory.
Hillel, D. 1980a. Fundamentals of Soil Physics. New York: Academic Press.
Hillel, D. 1980b. Applications of Soil Physics. New York: Academic Press.
Hiroma, T., and Y. Ota. 1990. Analysis of normal stress distribution under a wheel using a
viscoelastic model of soils. In Proc. 10th Intl. Conference of the Intl. Society for TerrainVehicle Systems, 241-252. Kobe, Japan.
Hopkins, M.A., J.T. Jenkins, and M.Y. Louge. 1992. On the structure of 3D shear flows. In
Advances in Micromechanics of Granular Materials, ed. H.H. Shen et al. Amsterdam: Elsevier.
Horn, R. 1981. Die Bedeutung der Aggregierung von Bden fr die mechanische Belastbarketi.
H10, ISBN: 37983
Horn, R. 1988. Compressibility of structured soils. In Interaction of Structured Soils with Water
and External Forces, ed. Drescher, Horn, and de Boodt. Catena Suppl. 11: 53-71.
Horn, R. 1990. Aggregate characterization as compared to soil bulk properties. Soil Tillage Res. 17:
265-289.
Horn, R., C.E. Johnson, H. Semmel, R. Schafer, and M. Lebert. 1992. Stress measurements in
undisturbed unsaturated soils with a stress state transducer (SST)-theory and first results. Z.
Pflanzenernaehr. Bodenk. 155: 269-274.
Horn, R., and M. Lebert. 1994. Soil compactibility and compressibility. In Soil Compaction in
Crop Production, ed. B.D. Soane and C. Van Ouwerkerk, 45-70. Amsterdam: Elsevier.
Johnson, C.E., and A.C. Bailey. 1990. A shearing strain model for cylindrical stress states. ASAE
Paper 90-1085. St. Joseph, Mich.: ASAE.
Johnson, C.E., and E.C. Burt. 1990. A method of predicting soil stress state under tires.
Transactions of the ASAE 33(3): 713-717.
Johnson, C.E., A.C. Bailey, and E. Cakir. 1992. Understanding soil response to multiple loading.
ASAE Paper No. 92-1051. St. Joseph, Mich.: ASAE.
Johnson, C.E., G. Murphy, W. Lovely, and R.L. Schafer. 1972. Identifying soil dynamic
parameters for soil machine systems. Transactions of the ASAE 15(1): 9-13.
Johnson, C.E., D.W. Wright, and A.C. Bailey. 1983. Compaction characteristics of some soil
mixtures. Transactions of the ASAE 26(5): 1337-1339.
Kezdi, A. 1976. Fragen der Bodenphysik. Duesseldorf, Germany: VDI-Verlag.
Khan, M.H., and D.L. Hoag. 1978. Three-dimensional stress-strain relationships of unsaturated
soil. ASAE Paper No. 78-1536. St. Joseph, Mich.: ASAE.
Kirby, J.M. 1989. Shear damage beneath agricultural tires: a theoretical study. J. Agric. Eng. Res.
44: 217-130.
Kondner, R.L. 1963. Hyperbolic stress-strain response: Cohesive soils. In Proc. Am. Soc. Civil
Engineers, J. Soil Mechan.. Foundations Div. 89(SM): 115-143.
Koolen, A.J., and H. Kuipers. 1983. Agricultural Soil Mechanics. Berlin: Springer-Verlag.
Koolen, A.J. 1994. Mechanics of soil compaction. In Soil Compaction in Crop Production, ed.
B.D. Soane and C. Van Ouwerkerk, 23-44. Amsterdam: Elsevier.
Kumar, L., and J.A. Weber. 1974. Compaction of unsaturated soil by different stress paths.
Transactions of the ASAE 17(6): 1064-1069.
Kushwaha, R.L., and J. Shen. 1994. The application of plasticity in soil constitutive modeling.
ASAE Paper No. 94-1072. St. Joseph, Mich.: ASAE.
Larson, W.E., S.C. Gupta, and R.A. Useche. 1980. Compression of agricultural soils from eight
soils orders. Soil Sci. Soc. Am. J. 44(3): 450-457.
Leeson, J.J., and D.J. Campbell. 1983. The variation of soil critical state parameters with water
content and it relevance to the compaction of two agricultural soils. J. Soil Sci. 34(1): 33-44.
Lowery, B., and R.T. Schuler. 1991. Temporal effects of subsoil compaction on soil strength and
plant growth. Soil Sci. Soc. Am. J. 55: 216-233.

Advances in Soil Dynamics, Vol. 2

Lyasko, M.J. 1982. The compaction action of agricultural tractors and machines on the soil and
methods of evaluation. Tractors and Agricultural Machines 10: 7-11.
Malvern, L. E. 1969. Introduction to the Mechanics of a Continuous Medium. Englewood Cliffs,
N.J.: Prentice-Hall.
Nichols, T.A., A.C. Bailey, C.E. Johnson, and R.D. Grisso. 1987. A stress state transducer for soil.
Transactions of the ASAE 30(5): 1237-1241.
Oida, A., and K. Yoshimura. 1980. Analysis of viscoelastic behavior of soil by means of finite
element method (II): application of FEM in two-dimensional stress state. J. Japan. Soc. Agric.
Mach. 41(4): 553-558.
Pearman, B.K. 1996. Front and rear tractor tire effects on soil stresses and compaction. M.S.
Thesis. Auburn University. Auburn, Ala.
Pearman, B.K., T.R. Way, C.E. Johnson, E.C. Burt, A.C. Bailey, and R.L. Raper. 1995. Soil
compaction from front and rear tires of an MFWD tractor. SAE Paper No. 95-2095, Society of
Automotive Engineers Intl. Off-Highway & Powerplant Congress and Exposition, Milwaukee,
Wisconsin, September 11-13, 1995. Warrendale, Pa.: SAE.
Pearman, B.K., T.R. Way, C.E. Johnson, E.C. Burt, A.C. Bailey, and R.L. Raper. 1996. Soil
stresses and rut depths from tires of a mechanical front wheel drive tractor. Transactions of the
ASAE 39(4): 1249-1257.
Perumpral, J.V., J.B. Liljedahl, and W.H. Perloff. 1971. The finite element method for predicting
stress distribution and soil compaction under tractive devices. Transactions of the ASAE 14(6):
1184-1188.
Petersen, C. T. 1993. Validation of two soil-compaction models for cylindrical stress states.
Transactions of the ASAE 36(5): 1293-1299.
Petersen, C. T. 1994. Modelling soil distortion during compaction for cylindrical stress load paths.
Eur. J. Soil Sci. 45: 117-126.
Pollock, D. Jr., J.V. Perumpral, and T. Kuppusamy. 1985. Finite element analysis of multipass
effects of vehicles on soil compaction. Transactions of the ASAE 29(1): 45-50.
Prevost, J.H. 1987. Modeling the behavior of geomaterials. In Geotechnical Modeling and
Applications, ed. S.M. Sayed. Houston, Tex.: Gulf Publishing.
Raper, R.L., A.C. Bailey, E.C. Burt, and C.E. Johnson. 1994. Prediction of soil stresses caused by
tire inflation pressures and dynamic loads. ASAE Paper No. 94-1547. St. Joseph, Mich.: ASAE.
Raper, R. L., P.W. Gassman, D.C. Erbach, and S.W. Malvern. 1987. Agricultural soil modeling
using finite-element analysis. ANSYS Conference Proc., Newport Beach, Calif.
Raper, R.L., C.E. Johnson, and A.C. Bailey. 1994. Coupling normal and shearing stresses to use in
finite element analysis of soil compaction. Transactions of the ASAE 37(5): 1417-1422.
Raper, R.L., C.E. Johnson, A.C. Bailey, E.C. Burt, and W.A. Block. 1995. Prediction of soil
stresses beneath a rigid wheel. J. Agric. Eng. Res. 61: 57-62.
Reece, A.R. 1977. Soil mechanics of agricultural soils. Soil Sci. 123(5): 332-337.
Roscoe, K.H., A.N. Schofield, and C.P. Wroth. 1958. On the yielding of soils. Gotechnique 9(8):
71-83.
Schafer, R.L., A.C. Bailey, C.E. Johnson, and R.L. Raper. 1991. A rationale for modeling soil
compaction behavior: an engineering mechanics approach. Transactions of the ASAE 34(4):
1609-1617.
Schafer, R.L., C.E. Johnson, A.J. Koolen, S.C. Gupta, and R. Horn. 1992. Future research needs in
soil compaction. Transactions of the ASAE 35(6): 1761-1770.
Schofield, A., and P. Wroth. 1968. Critical State Soil Mechanics. McGraw-Hill, London.
Schuler, R.T., and R.K. Wood. 1992. Soil Compaction. In Conservation Tillage Systems and
Management, MWPS-45. Iowa State University, Ames, Iowa: Midwest Plan Service.
Shen, J., and R.L. Kushwaha. 1995. Stress-strain rate relation of agricultural soils. Can. Agric. Eng.
37(1): 19-28.
Smith, D.L.O. 1984. A simple prediction model for soil compaction under various wheel loads and
geometries as an aid to vehicle design. In Proc. 8th Intl. Conference of the Intl. Society for
Terrain-Vehicle Systems 737-749. Cambridge, England.

Advances in Soil Dynamics, Vol. 2

Soane, B.D. 1985. Traction and transport systems as related to cropping systems. In Proc. Intl.
Conference on Soil Dynamics 5: 863-935. Auburn, Ala.: National Soil Dynamics Laboratory.
Soane, B.D., and C. van Ouwerkerk (eds.). 1994. Soil Compaction in Crop Production.
Amsterdam: Elsevier.
Shne, W. 1958. Fundamentals of pressure distribution and soil compaction under tractor tires.
Agricultural Engineering 39: 276-281, 29.
Shne, W., W.J. Chancellor, and R.H. Schmidt. 1962. Soil deformation and compaction during
piston sinkage. Transactions of the ASAE 5(2): 235-239.
Srivastava, A K., and J.F. Steffe. 1987. The rheological approach to soil modeling. Transactions of
the ASAE 30(6): 1661-1672.
Taylor, J.H., and W.R. Gill. 1984. Soil compaction: state-of-the-art report. J. Terramechanics
21(2): 195-213.
Taylor, J.H., A.C. Trouse, Jr., E.C. Burt, and A.C. Bailey. 1982. Multipass behavior of a pneumatic
tire in tilled soil. Transactions of the ASAE 25(5): 1229-1231, 1236.
Telles, J.C.F. 1983. The Boundary Element Method Applied to Inelastic Problems. Berlin:
Springer-Verlag.
Ting, J.M., B.T. Corkum, C.R. Kauffman, and C. Greco. 1989. Discrete numerical model for soil
mechanics. J. Geotech. Eng. 115(3): 379-398.
Upadhyaya, S.K. 1994. Soil constitutive modeling-linear and nonlinear elasticity: A tutorial
presentation. ASAE Paper No. 94-1071. St. Joseph, Mich.: ASAE.
Van den Akker, J.J.H., and A.L.M. Van Wijk. 1985. A model to predict subsoil compaction due to
field traffic. In: Soil compaction and regeneration. In Proc. Workshop on Soil Compaction,
Consequences and Structural Regeneration Processes, ed. G. Monnier and J.J. Goss, 69-84.
Avignon, France, 17-18 September. Rotterdam, Netherlands: A.A. Balkema.
Vanden Berg, G.E., W.F. Buchele, and L.E. Malvern. 1958. Application of continuum mechanics
to soil compaction. Transactions of the ASAE 1(1): 24-27.
Vanden Berg, G.E., and W.R. Gill. 1962. Pressure distribution between a smooth tire and the soil.
Transactions of the ASAE 5(2): 105-107.
Venturini, W. S. 1983. Boundary Element Method in Geomechanics. New York: Springer-Verlag.
von Estorff, O., and E. Kausel. 1989. Coupling of boundary and finite elements for soil-structure
interaction problems. Earthquake Engineering and Structural Dynamics 18: 1065-1075.
Wood, R.K. 1988. Thrust-dynamic load relationships for an agricultural tire. Ph.D. dissertation.
Auburn University, Auburn, Ala.
Wood, R.K., R.C. Reeder, M.T. Morgan, and R.G. Holmes. 1993. Soil physical properties as
affected by grain cart traffic. Transactions of the ASAE 36(1): 11-15.
Wroth, C.P., and G.T. Houlsby. 1980. A critical state model for predicting the behavior of clays. In
Proc. Workshop on Limit Equilibrium, Plasticity and Generalized Stress-strain in Geotechnical
Engineering, 592-627. New York: ASCE.
Wulfsohn, D. 1994. Critical state soil mechanics. ASAE Paper No. 94-1073. St. Joseph, Mich.:
ASAE.
Wulfsohn, D., B. A. Adams, and D. G. Fredlund. 1996. Application of unsaturated soil mechanics
for agricultural conditions. Can. Agric. Eng. 38(3): 173-181.
Wulfsohn, D., B. A. Adams, and D. G. Fredlund. 1998. Triaxial testing of unsaturated agricultural
soils. J. Agric. Eng. Res. 69(4): 317-330.
Yamamoto, Y. 1972. Particle model methods for stress analysis of soil (I): on the static particle
model system and numerical analysis. J. Japan. Soc. of Agric. Mach. 34(4): 318-326.
Yamamoto, Y. 1975. Particle model methods for stress analysis of soil (II): a numerical method for
unconstrained nonlinear optimization problems. J. Japan. Soc. of Agric. Mach. 37(3): 309-314.
Yong, R.N. 1984. Track-soil interaction. J. Terramechanics 21(2): 133-152.
Yong, R.N., and E.A. Fattah. 1976. Prediction of wheel-soil interaction and performance using the
finite element method. J. Terramechanics 13(4): 227-240.
Yong, R.N., and B. Warkentin. 1966. An Introduction to Soil Behavior. New York: Macmillan.
Yong, R.N., and B. Warkentin. 1975. Soil Properties and Behaviour. Amsterdam: Elsevier.

Vous aimerez peut-être aussi