Soil Dynamics in Tillage and Traction: Agriculture Handbook No. 316
Soil Dynamics in Tillage and Traction: Agriculture Handbook No. 316
Soil Dynamics in Tillage and Traction: Agriculture Handbook No. 316
SOIL DYNAMICS
in
TILLAGE AND TRACTION
MBP0000263
SOIL DYNAMICS
in
TILLAGE AND TRACTION
By
WILLIAM R. GILL
Soil Scientist
Agricultural Engineering Research Division
and
Soil and Water Conservation Research Division
1. INTRODUCTION 1
1.1 Historical 1
1.2 Soil Dynamics 3
1.3 Research Centers 5
1.3.1 The National Tillage Machinery
I^aboratory, Auburn, Ala. 6
1.3.2 The Army Mobility Research Center,
Vicksburg, Miss. 7
1.3.3 The Land Locomotion Laboratory,
Warren, Mich. 8
1.3.4 The National Institute of Agricultural
Engineering, Silsoe, England 8
1.3.5 Institute of Fundamental Research in
Agricultural Engineering, Volkenrode, Germany 10
1.3.6 Institute for Agricultural Mechanization,
Konosu, Japan 11
1.3.7 Other Research Centers 11
2. DYNAMIC PROPERTIES OF SOILS 14
2.1 Introduction 14
2.2 Stress in Soil 14
2.3 Strain in Soil 17
2.4 Stress-Strain Relations __ _ Z_~ZI 20
2.5 Soil Strength '___ 22
2.6 Stress Distribution _~II_ II_ 23
2.7 Strain Distribution _ _ _ _ _~~~___I"_" ..29
2.8 Yield in Soil __ I" __"_'_ __ "_""!""" 31
2.8.1 Shear IIII—II-"_-I_-I__II_I_._"_I 32
2.8.2 Compression __ 36
2.8.3 Tension _ _ _ ~~ ___ 37
2.8.4 Plastic Flow I_ _""" _ 39
2.9 Rigid Body Soil Movement I__I_ZIII_I 39
2.9.1 Momentum ~~ "~ Z"Jl ~ 40
2.9.2 Friction IIIIIIII_IIII__I__ _ 40
2.9.3 Adhesion IIIIIII_I_ZI_"~II_I 42
2.9.4 Abrasion II_I~I~II II __"_ II 52
2.10 Dynamic Versus Static Properties ___I__I_I_ I_ I "_"_" _"_I_ 53
3. ASSESSMENT OF THE DYNAMIC PROPERTIES OF SOIL" "___ 55
3.1 Soil as a Physical System _ _ 55
3.2 Dynamic Parameters I" I I H "I QQ
3.2.1 Measuring Independent Parameters _I __~ _ _ ~~ 65
3.2.1.1 Shear __II I_~II aí
3.2.1.2 Tension IIIIIII Z.Jl.JlJl^ 74
3.2.1.3 Compression __ __ _ _ CA
3.2.1.4 Plastic Flow __"_ öQ
3.2.1.5 Friction I_ 22
3.2.1.6 Adhesion I-I-I'III _"~_'r_~~rri 89
3.2.2 Measuring Composite Parameters QQ
3.2.2.1 Penetration _ _ _ 04
3.2.2.2. Bearing Strength
3.2.2.3 Induced Strength ___ _ __ _ _ 100
3.3 Measuring Gross Dynamic Behavior 1X9
3.3.1 Rupture ~""~_ IAQ
3.3.2 Blast Erosion i}t¿
3.3.3 Abrasion I ,XQ
3.3.4 Movement -"""IIIHIIIIIIIIIIIIIII—I—I 113
VI
CONTENTS Vil
1.1 Historical
Soil dynamics, a phase of soil science and mechanics concerned
with soiîs in motion, may be defined as the relation between forces
applied to the soil and the resultant soil reaction. This definition does
not restrict the source of the force applied to the soil ; consequently,
the dynamic reactions that result from the natural forces of wind,
water, and other sources are included. Reactions due to wind and
water are of paramount importance in erosion and hydrology and
the mechanics of these reactions are being studied ( 72, 391 )}
Only reactions caused by mechanical forces applied directly to the
soil are considered here. The dynamic reactions of soil in tillage and
traction affect the design and use of machines that handle soil.
Because the interrelations are of primary interest, ihç^ tool (or trac-
tion device) and the soil must be considered together.
In order not to restrict the final application of the research find-
ings, tillage is defined as mechanical manipulation of the soil (for
any purpose). This definition may be difficult to accept, but the
same type of plow that plows soil for agricultural purposes may also
be the basic tillage unit of a machine that lays antitank mines for
military purposes. The bulldozer does exactly the same work when
leveling land for irrigation and drainage purposes as when leveling
land for new buildings, roads, and parks. Tillage is therefore defined
in terms of applying forces rather than in terms of the reason for
which the forces are applied. .
Traction is the force derived from the soil to pull a load. This torce
is exerted against the soil by a traction device such as a wheel, track,
winch sprag, or spade. The dynamic resistance of the soil to provide
traction is supplied through an interaction between the traction de-
vice and the soil. This interaction is very complex and little head-
way has been made in solving some of the problems that result from
the interaction.
The practical importance of soil dynamics in tillage and traction
has been known for many years; however, soil dynamics research
has been conducted only since 1920. Until that time no concerted
effort was made to bring information together and to direct research
along fundamental lines so as to solve complex problems in tillage
and traction. What may have been the first doctorate thesis in agri-
cultural engineering in the United States was written by E. A.
White at Cornell University in 1918. It was entitled "A Study of the
Plow Bottom and its Action Upon the Furrow Slice." No doubt this
thesis will stand as í\\Q landmark identifying the point where a more
theoretical approach in the study of tillage tools began.
The inability to measure and characterize soils has been the limiting
factor in soil dynamics.
The factors suggested by Nichols may presumably be related to
each other in some quantitative fashion. Soil dynamics constitutes
the basis on which to establish, interpret, and use all of the possible
relations that are involved. The problem may be broken down into
two general areas that center around practical objectives: (1) a
basic area, in which soil-machine dynamic relations and interactions
are established; (2) an application area, in which the performance
of a machine in a soil system is determined. The soil-machine me-
chanics should apply to all possible soil reactions and serve as a
basis on which the dynamic properties of soil can be used. Evalua-
tion of performance cannot be restricted to dynamic properties since
economic, soil, machine, plant, social, and other factors must be
considered. The number and kind of factors to be considered depend
on the specific application.
1.3 Research Centers
Following the decline in research in soil dynamics during World
War II, the resumption of research was slow. Since 1950, however,
the work has been expanding at an increasing rate and new research
centers are being developed. Because of the specific requirements of
soil dynamics research, future work will probably be conducted at
these research centers where specialized research facilities and per-
sonnel can be concentrated. For example, results obtained from
models are difficult to apply to full-scale equipment. Hence, appa-
ratus is needed that can handle full-size equipment and large
volumes of soil. The soil at these research centers is usually selected
on the basis of its physical properties, and it is frequently moved
to the center from great distances.
Rarely are attempts made to reconstruct soil profiles; rather, uni-
form soil conditions are established to provide a homogeneous soil
material. Special soil-fitting machinerv is required to establish and
maintain specific soil conditions {351). Considerable improvement
in this type of machinery is needed.
Another requirement arises since accurate measurements can be
made only when the device being examined is moved through the
soil at controlled speeds, orientations, and depths while resulting
forces are being measured. A track or guidance system, therefore,
must be used to maintain a fixed path for the device throughout a
test lane. Attempts to use portable tracks have not been successful,
and more workers are resorting to fixed tracks. The fixed-track sys-
tem is so specialized that it has little other use; hence, it must be
used continuously for that purpose once it is constructed. Because of
these factors, the facilities required in conducting the research are
not only specialized but also very expensive. The result is that these
facilities are constructed only when a long-range specialized program
is to be pursued. Since no two of the existing research centers have
facilities with the same capabilities, research on all types of prob-
lems cannot be conducted at every center. Specific problems may be
handled better at one center than another. However, the willingness
of the research agencies to cooperate to prevent needless duplication
is most encouraging.
6 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
various forces acting on wheels or tools and the soil can be studied
to determine the basic dynamic relationships and soil reactions. The
soils vary in mechanical composition from sand to clay and provide
SOIL DYNAMICS IN TILLAGE AND TRACTION <
a wide range of soil material. The soils are prepared for tlie various
investigations with specialized machinery, and moisture can be con-
trolled to some degree by covering the plots or by watering them \yith
a special mobile sprinkler apparatus. Indoor facilities provide bet-
ter control of the environment for two large bins. Indoor labora-
tories provide facilities for emphasis on studies that do not require
full-scale tillage and traction machinery.
1.3.2 The Army Mobility Research Center, Vicksburg, Miss.
The Army Älobility Research Center, U.S. Army Engineers "Water-
ways Experiment Station, conducts research to determine the traffic-
ability of soil. Major emphasis is placed on soft soils where im-
mobilization of vehicles or aircraft is likely to occur.
A unique feature of the Center is an indoor small testing facility.
As shown in figure 2, this facility utilizes an overhead set of rails
FIGURE 2.—Movable soil bins and dynamometer, at the Army Mobility Research
Center, for determining the performance of normal-size wheels on different
soils and soil conditions. Electrical signals generated by the detecting device
are transmitted through instrumentation cables to recording equipment.
(Photograph courtesy of U.S. Army Engineers Waterways Experiment
Station)
on which the test car is operated. The soil bins are fitted with the
desired soil and then positioned under the guide rails during the
test. Forces can be measured on wheels at speeds up to 25 m.p.h.
A large indoor soil crushing and mixing plant is available for pre-
paring soil so that measurements can be made on a continuous basis
at any time of the year.
This Center probably has the largest concentration of (1) person-
nel; (2) instruments for measuring, recording, and processing re-
search data; and (3) supporting soil evaluation facilities for soil
8 AGRICULTURE HANDBOOK 316, U.S. DEPT. Oí' AGRICULTURE
FIGURE 3.—A small model vehicle at the Land Locomotion Laboratory provides
an economical means of studying soil and vehicle parameters. The char-
acteristics believed to be of importance in land locomotion are being studied
for many soil conditions. (Photograph courtesy of U.S. Army Ordnance
Corps Land Locomotion Laboratory)
FIGURE 4.—A mobile dynamometer unit .it the National Institute of Agricul-
tural Knglneerlng used to study the forces on tillage tools under dynamic
operating conditions. The electronic signals produced by the measuring
devices are recorded in an Instrument vehicle, which drives beside the test
vehicle. (Photograph courtesy of National Institute of Agricultural En-
gineering )
FIGURE 6.—The mobile plow testing unit used by the Institute of Fundamental
Research in Agricultural Engineering is shown with a disk plow in operating
position. An auxiliary engine permits rotation of the disk during operation.
(Photograph courtesy of Institute of Fundamental Research in Agricultural
Engineering)
FIGURE 8.-—A highly mechanized model testing apparatus used by the Cater-
piller Tractor Company. (Photo courtesy of Caterpillar Tractor Company)
2.1 Introduction
The dynamic properties of soil are properties made manifest
through movement of the soil. If a block of soil resting on a flat
surface is moved, the resultant friction is a dynamic property of the
soil; this property cannot be determined until movement occurs.
Similarly, as loose soil is compacted, its strength increases; hence,
strength is a dynamic property of soil. When soil moves, forces act
that cause deformation or actual physical displacement. To describe
the relations between the applied forces and resulting deformation,
certain basic mathematical equations containing parameters are re-
quired. The parameters are measures of the dynamic properties of
the soil.
The structure or texture of soil is not a dynamic property. The
structure may be changed as a result of movement, but it may be
measured both before and after the movement. Such is not the case
for dynamic properties. Studying dynamic reactions is difficult be-
cause the physical measurements must be made during the action.
In addition, the insertion of measuring equipment into the soil mass
may affect the soil reaction. The equipment may behave differently
than the soil, if, for example, it is harder or softer than the soil. In
spite of these difficulties, considerable progress has been made both
in identifying dynamic properties and in utilizing these properties
to describe the reaction of soil to forces {16).
2.2 Stress in Soil
Describing forces acting in soil is not as simple as first appearances
indicate. When discussing a finite block of soil, as in the case of
sliding friction, vector representation of the gravity forces, me-
chanically applied forces, and friction forces is probably sufficient.
Even describing the distribution of forces within the finite block is
not too difficult. When considering, however, that the mass of soil
under discussion is usually a semi-infinite medium where the applied
forces are distributed over a small finite section of the boundary, the
problem becomes much more difficult. The concept of force per unit
area becomes meaningless in a three-dimensional semi-infinite me-
dium where neither direction nor a finite area is fixed. A method
IS thus required to describe the forces acting at each point within
and on the medium.
The state of stress at a point is one method of describing forces
within a medium. The method can be developed in a rigorous math-
ematical manner if the material is assumed to be continuous—that is,
without any holes or gaps, unfortunately, the soil is not continuous
since it has pores and is a granular material. The mechanics of the
14
SOIL DYNAMICS IN TILLAGE AND TRACTION 15
For the limit to exist and have physical meaning, the area must be
continuous ; hence, the mathematical requirement for the continuum.
The stress vector T is usually broken into components normal to the
plane and tangent to the plane. Although many methods are used
to designate stresses, one common method uses the Greek letter sigma
(er) for normal stresses and the Greek letter tau (r) for shearing
stresses that are tangent to the plane. An appropriate subscript is
generally used to indicate the plane with which each stress is as-
sociated.
If another plane is passed through the same point O, usually a
different stress vector will be found. Since an infinite number of
planes can be passed through the same point some simple method is
required for calculating the stress vector on any plane once certain
specified values are given. It is possible to proceed along well-known
lines established by the mechanics of a continuous medium ( 119,
190 ). Specifying the stress vectors on three mutually perpendicular
planes is sufficient. Vanden Berg ( ^58 ) has pointed out that such
a specification requires nine quantities and in this case represents an
entity that is, by mathematical definition, a tensor. These nine com-
ponents of a stress tensor at point O are shown in figure 10. Since
graphically visualizing three intersecting planes is difficult, the
normal representation embodies a small cubical volume element.
The element is oriented with respect to a triad of orthogonal axes so
that the sides of the cube are perpendicular to the axes. The com-
ponents of stress shown are considered to be those that will be
present when the volume of the element shrinks to zero at point O.
16 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
z i<^Z
^ ZX
^
^
^z
."^v
J^ ^v ^YX
'<y^ )
^
^
To correctly describe the forces acting on and within soil, the nine
values must be specified at each point including the boundaries.
That the entity of nine values forms a tensor is in reality helpful,
since the powerful techniques of tensor analysis become available to
assist in solving the problem.
Certain simplifications are possible. From symmetry and equi-
librium, it can be shown that TX '7'xz — '7'z: and Tv By
utilizing these identities, three of the unknowns can be eliminated so
that only six independent values must be specified to describe the
state of stress at one point in the soil.
A property of the stress tensor is that the coordinate axes can
always be rotated (and with them the imaginary planes bounding
the cubical volume element) so that all shear stresses will be zero
and only normal stresses will act on the three mutually perpendicular
planes. The three normal stresses are the well-known principal
stresses, and their directions are the principal axes of the stress state.
The mean normal stress cr^ is defined as
(Tm — 1/3 (cri + a-2 +0-3), (2)
where CTI, 0-2, and era are principal stresses. The quantity is an in-
variant of the stress tensor so that the following relationship is also
true
o-^ = 1/3 (0-^ + 0-2,+o-z), (3)
where (TX, o-y, and o-^, are the normal stresses referred to any coordin-
ate axes of the stress state. A spherical state of stress is often dis-
cussed where each normal stress is equal to the mean normal stress—
that is, similar to the stress state in water at rest. Subtracting the
spherical stress from the total stress by appropriate mathematical
methods leaves a deviatoric stress. The total stress state is thus
mathematically separated into two components—spherical and devia-
SOIL DYNAMICS IN TILLAGE AND TRACTION IT
d€ = f-, (5)
where de and dl are differential values of e and I as defined above.
Another measure of deformation is the change in angle between
two initially mutually perpendicular line elements. The usual defini-
tion (fig. 11) is
y = tan (90°-i//), (6)
where y = shearing strain,
90° — i// = deformation angle.
As in stress, six independent values must be specified to define
strain. Three mutually perpendicular longitudinal strains and three
shearing strains in the planes formed by the triad of longitudinal
strains are the required values. From these values, the longitudinal
strain of any line element or the shearing strain of any two initially
mutually perpendicular line elements radiating from the point can
be calculated. These six independent values form a second order
18 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
/■
I
I
I
I
I
I
I
^-^P'
A
FIGURE 11.—Shearing strain defined as change in 90° — i//, the angle
between two line elements.
= In -y-. (8)
B. C^ ^D
I
I
-5L- A- -.5L-
I
I I
._L. .J
scribe soil behavior has been delayed because of the wide range of
soil and its behavior. The observed nonlinearity of stress-strain
relations has yet to be mathematically described. Simplifying as-
sumptions have been made to circumvent stress-strain requirements
so that some degree of order can be made of soil reaction to forces.
These approaches are discussed in detail in chapter 3. One must
remember, however, that any approach not based on accurate stress-
strain relations can only approximately represent the true behavior
of soil. The development of suitable stress-strain relations, which
in turn define parameters, is one of the more important areas of re-
search in soil dynamics. Until such relations are determined, one
cannot even study the dynamic properties of soil because they have
not been clearly identified.
2.5 Soil Strength
Soil strength is the ability or capacity of a particular soil in a
particular condition to resist or endure an applied force. Soil
strength might also be defined as the capacity of soil to withstand
deformation or strain since strength is not evident without strain.
The concept of soil strength is thus quite clear and easily understood.
Soil strength is a physical quantity.
The problem is to measure and describe strength so that a definite
series of numerical values can be assigned. This problem has not
been satisfactorily solved to date. The wide range of strengths ob-
served in soils is one difficulty ; another is that the strength actually
changes when force is applied and movement occurs. Changing
strength is exhibited in many other materials but to a much lesser
degree than in soil. Strength is thus truly a dynamic property of
soil.
One obvious way of describing soil strength is to use the para-
meters of stress-strain equations. The numerical values of the para-
meters for each condition would quantitatively describe its strength.
Viscoelastic, yiscoplastic, and fluid mechanics equations are examples
of stress-strain equations.
Another way of describing soil strength is to evaluate the para-
meters involved in yield conditions. Yield conditions are fully dis-
cussed later in this chapter (sec. 2.8). Yield parameters are different
from the two parameters, modulus of elasticity and Poisson's ratio,
that appear in elasticity stress-strain equations. Accurate evaluation
of the parameters of stress-strain equations and of yield conditions
for soil provides a logical and sufficient means for describing strength.
Lack of both stress-strain equations and adequate yield conditions for
soil has prompted the use of simulation-type tests to describe soil
strength. Some examples of these tests are discussed in chapter 3.
The inability to describe soil strength accurately has dictated the
kind of much of the experimental work dealing with soil dynamics.
For example, tillage tools or traction devices can be compared only
in the same soil conditions. Results obtained in one soil condition
cannot be compared with those obtained in another soil condition.
This is true even when the same soil type is used, because the strength
may be different. The recent use of artificial soils ( 229, 2J4S, Slß^
383 ) is an attempt to eliminate strength changes, so that a variable
SOIL DYNAMICS IN TILLAGE AND TRACTION 23
(A)
(B)
vP (14)
(Tn = cos"""^ (^ sin^ <^,
^TTT^
o-t = 0,
vP (15)
Th = r^ cos''"^ (f) sin (^,
where v is the concentration factor and the other terms are as indi-
cated in figure 14. Eotating the volume element about cTt axis so CTn
f/ù,^:^///Jk^/??
i_z
FLUID CHAMBER SENSING DIAPHRAGMV
•STRAIN GAGES'
(A) (B>
CELL
CELL
(A) (B)
than the soil, a lower value will be obtained since the load will be
transferred from the transducer to the surrounding soil, as shown in
figure 17, B. Thus the size, shape, and resiliency that the transducer
presents to the soil mass will affect the transducer's reading.
Within certain limits, the transducers appear to give valid read-
ings. A diameter-thickness ratio of 5:1 appears to give satisfactory
results for circular transducers, although ratios as low as 2:1 have
been used. Although no definite evidence has been presented, the
exact ratio is probably a function of the degree of soil settling dur-
ing loading and the closeness of the transducer to some firm object or
layer in the soil. No transducer has been devised that can make
measurements without some degree of disturbance to the soil, but use
of transducers in snow ( 4^0 ) shows that the degree of disturbance
may not preclude their profitable and reliable use until better trans-
ducers are devised. Improved instrumentation will permit studies
of stress distributions that heretofore could be evaluated only from
empirical calculations.
Certain stress indicators may be used to qualitatively determine
stress distribution in soil. Since compaction is caused by stress, the
density of soil increases as stress increases. The location of com-
pacted areas after loading a homogeneous soil is, therefore, an indi-
cation of a stress concentration. In snow studies, the "Nakaya Pit
Burning" technique ( 309^ 4^5^ JßO ), where smoke from an oil fire
deposits more carbon on denser snow, employs this principle so that
compacted areas can be detected. Materials that require small
SOIL DYNAMICS IN TILLAGE AND TRACTION 29
strains for fracture—that is, stress coats—can be used in only rare
cases. Similarly, pliotoelastic techniques have highly limited ap-
plications in soil because of the extreme porosity of the soil, the in-
fluence of the material on soil strenjith, and the large strains result-
ing from yield. A soil box with one side made of thick glass plate
has been used at the National Tillage Machinery Laboratory to locate
stress concentrations. Reaction of the soil to a tool can be viewed
through the glass side. The tool is moved through the soil im-
mediately parallel to and against the glass, and the compacted areas
are located visually. The stress concentrations are obvious in some
soils, as shown in figure 18.
(A) (B)
the proposed path of the vehicle so that the soil was exposed in a
small area under the path. A movie camera in the pit recorded the
displacement of a marker in the soil as the vehicle passed over the
site.
Strain may also be measured on a volumetric basis under dynamic
loadings. Hovanesian ( 185 ) devised a volumeter utilizing a bal-
loon filled with soil, which was buried in a larger mass of the same
soil. The balloon opening was connected to a horizontal pipette
containing a drop of mercury that acted as a rolling seal. The soil
inside the balloon was assumed to act in the same manner as the sur-
rounding soil, so that it represented a small volume element of soil.
Any volume change inside the balloon forced air from the balloon
into the pipette and displaced the mercury seal by an amount equal
to the volume change. If the initial volume of soil inside the balloon
is not known, the final volume can be determined by water dis-
placement after the volumeter is removed from the soil. A slight
vacuum applied to the balloon insures that the final configuration of
SOIL DYNAMICS IN TILLAGE AND TRACTION 31
the compacted soil is maintained. The volumeter is sensitive to
temperature, however, so that it must be used under isothermal con-
ditions. The volumeter was later modified { 186 ) to obtain an elec-
tric signal that could be recorded.
Figure 20 shows the components of the original volumeter as well
CO
(O
UJ
STRAIN (€)
desired area to describe the stress state. Note that if any two of the
principal stresses are equal, the three circles become one and the area
becomes the circumference of one circle ; the one circle is Mohr's rep-
resentation in two dimensions. Maximum shearing stress, largest
principal stress, and other stresses can be easily obtained from the
graphical representation of the state of stress.
Failure by shear has a clear meaning for brittle materials. The
classical example is the brittle-type shear fracture that develops
when a solid cylinder of material is loaded in compression as shown
in figure 23. This type of failure is also observed in soil. A similar
type of failure is observed in some soil conditions where no definite
fracture occurs but where the diameter of the specimen under test
gradually increases with load. The stress state that causes fracture
or incipient plastic now—that is, larger diameter and hence perma-
nent deformation—is a measure of shear failure. The definition of
failure by shear is thus some function of the stress state that just
causes the failure. In soil, both complete fracture and incipient
plastic flow have generally been indicated by the same stress function.
A review of shear by Jaeger ( 190 ) indicates that the first theory
34 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
FIGURE 24.—A Mohr envelope of stresses from which soil parameters TO and 0
can be determined.
_>(fl*iin?wT.
total area (air, solid material, and water), the area of solid material
and water, and the area of only solid material have all been used for
this purpose. Using any area other than total area is an attempt to
incorporate behavior explained from a granular model into the
mathematical model of the continuum. As pointed out earlier, this
technique is not inconsistent and does not violate either model. The
area to use should be the one that results in the best representation
of tension failure.
The Mohr envelope of stresses (fig. 24) also indicates tensile
stresses. Along the ordinate, stresses to the left of the abscissas are
tensile stresses and those to the right are compressive. Willetts
( 50p ) has constructed the locus of shearing stress r from a combi-
nation of measured values of tensile and compressive stress loadings
that were imposed to cause soil failure. When uniaxial tensile stress
is applied to cause soil failure in a rigid body system, there should
be no shearing stress. When soil does not act as a rigid body, the
behavior induces shearing stresses. Accordingly, a shearing stress r
will accompany each tensile stress cr. Vomocil and Waldron ( Ii.68 )
have studied Mohr relations in which the shearing stress was less
than To^ In this case, the magnitude of the shearing stress may be
calculated from the ratio of the negative major to minor principal
stresses and the angle of internal friction.
Direct tension is seldom applied to soil by a tillage tool or traction
device. Tension failure does have physical significance, however,
and may sometimes be induced during soil manipulation. Although
soil is often thought to be incapable of sustaining a tensile force, its
tensile strength may be very high. Table 2 gives tensile failure
strengths of several clay and loam soils. The soils were dry and
cemented; their strength decreased rapidly on wetting. Neverthe-
less, the values indicate that soil can be strong in tension.
The.shear criterion for failure (equation 18) implies that soil
strength is composed of two factors—cohesion and friction. Cohe-
sion is defined as the force that holds two particles of the same type
together. When soil fails in tension, only the cohesive part con-
tributes to resistance since the normal stress is zero. Such reasoning
immediately suggests that tensile failure may be exactly the measure
of cohesion. Unfortunately, no one has yet devised a means of meas-
uring direct shear at zero normal stress so the question has never
been resolved. In reality the argument is of academic interest only
SOIL DYNAMICS IN TILLAGE AND TRACTION 39
scouring moldboard plow involves both rigid and nonrigid body de-
scriptions; often the same mass of soil is involved in both descrip-
tions. This double involvement is complicated but justified since we
are using different mathematical mode^ls to represent only the ob-
served behavior.
2.9.1 Momentum
Momentum does not involve a dynamic property of the soil but is
often involved in the action of tillage and traction equipment. Mo-
mentum is the product of mass and velocity. Since a rigid block of
soil has a definite mass and when moving has a definite velocity, its
momentum can be defined. If properly utilized, this momentum can
have considerable inñuence on soil breakup. Besides describing the
momentum of a rigid block of soil, the Newtonian laws of motion
rigorously represent such factors as forces, accelerations, velocities,
displacements, and inertias. Thus, a mechanics of detached soil
masses is available and is clearly defined. The inherent assumption
when using Newtonian laws is that the mass being considered behaves
as a rigid body.
2.9.2 Friction
When two rigid bodies of soil move with respect to each other,
forces act on the mutual contact surface. The general laws of
Coulomb friction apply, but the exact nature of frictional forces is
not yet known {80). The usual procedure is to separate the acting
forces into those normal to the surface and those tangent to the
surface.
Following Coulomb's concept, a coefficient of friction can be defined
F
/^ -^ = tan i/;, (19)
FIGURE 26.—Normal force and friction force between two rigid bodies of soil.
V///////
FIGURE 27.—A method for inducing direct shear failure along a predetermined
surface.
soil is transmitted through the soil. Shims are placed between the
upper and lower parts of the box during filling. Eemoving the
shims for measurement assures the necessary clearance. Failure oc-
curs between the two halves of the box so that F in figure 27 is
equivalent to the frictional force. If the coefficient of friction is
considered to be independent of normal load, a series of different
normal loads plotted versus the frictional force gives a straight line.
The slope of the line is a measure of the coefficient of friction /x.
Nichols ( SIS ) and others ( 12 ) have measured shear failure on two
surfaces as the central portion of a cylinder is moved. There appears
42 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
the radius of the tube r, the angle of contact a, and the theoretical
maximum tube size R. The theoretical radius R is determined by the
spherical section of the meniscus, which in turn is determined by
the angle of contact, and R is interpreted as the maximum radius
tube that a given surface tension could support ; that is, r can theo-
retically be increased to i?.
The geometry in figure 29 shows that
cos a = (21)
R'
and substitution of equation 21 into equation 18 yields
h =
2r (22)
pgR
where R is now the radius of the theoretical limiting tube size or
pore size that a given surface tension will support. Since the term
A p ^ is a pressure term and gives a measure of pressure, equation 22
can be written
P -^ (23)
^ ~~ R'
where P is now pressure in the continuous solution. Since h in
equation 20 represents a pressure deficiency—that is, less than atmos-
pheric pressure—F in equation 23 is the basis for using moisture
tension measurements as a means of measuring pore size distribution
in soil. As the moisture tension is increased by some means such as
a moisture tension table ( 192 ) or a pressure membrane device
44 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
( JfJfS ), the soil pores will support the tension until the capillary
force is overcome and air enters the pore and causes drainage. Equa-
tion 23 gives the relationship between the applied moisture tension
and the theoretical maximum radius of water-filled pores. By using
equation 23 and the volume of soil solution removed under suction,
the size and amount of pore space in a soil sample can be determined.
Since the pores in soil are irregular in shape and vary in size, the R
calculated in equation 23 is a value that is theoretically equivalent to
the actual soil pore.
McFarlane and Tabor ( 26^^ 265 ) demonstrated the importance of
surface tension on adhesion by measuring the adhesion between a
glass sphere and a glass plate under saturated conditions. Figure 30
shows the water film between a flat plate and a sphere at saturation.
Since both surfaces in the experiment were glass, the contact angle
is the same. At saturation the water film contacts the plate at a
circle whose diameter is 4Ä, where R is the radius of the sphere.
Following the concept of capillary rise, the force F between the
sphere and plate is equal to the perimeter of contact multiplied by
the surface tension T^ and the cosine of the angle of contact a. This
relation is expressed as
F = 4:7rRT cos a. (24)
McFarlane and Tabor measured F, the adhesion force, for various
radii of beads. They used a clever method where the spherical glass
beads were suspended on the end of a fine fiber. The vertical plate
and bead were brought into contact; then the plate was moved back
until the bead fell away under the action of gravity. Figure 31 shows
the forces acting and shows how the angle 0 is a measure of F. Ad-
hesions as low as 10"^ grams could be readily measured with the
apparatus. Figure 32 shoAvs the results of this work. From the
slope of the line and equation 24, the surface tension for water could
be calculated. This method resulted in a value of 67.3 dynes per
centimeter, whereas the accepted surface tension of water at the same
temperature is 72.7 dynes per centimeter. They felt, however, that
the value was close enough to demonstrate the importance of surface
SOIL DYNAMICS IN TILLAGE AND TRACTION 45
FIGURE 31.—A method used to determine the force of adhesion between a glass
bead and a flat plate.
100 r
tension on adhesion. The values for surface tension for other liquids
in table 3 clearly verify the effect of surface tension on adhesion.
Fountaine ( 127 ) demonstrated the effect of moisture tension on
adhesion. He reasoned that when a soil is saturated and a continu-
ous water film exists between a plate and soil, tension m the \yater at
equilibrium would come to the same value throughout the soil mass.
To achieve equal tension most of the water-air menisci will move
into channels whose dimensions correspond with the tension governed
by the relation expressed in equation 23. Some of the menisci may
have to assume "forced" curvatures, since they are unable to reach
46 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
SINTERED
GLASS BRASS
SOIL- PLATE
WATER
SOIL- ^SOIL
SCRAPED AWAY
COTTON- (B)
WOOL
WEIGHT
FIGURE 33.—^An apparatus to measure the normal force between soil and a
metal plate. (Fountaine, Jour. Soil Sei. {121),)
8 12 16 20 4 8 12 16 20 4 8 12 16 20
MOISTURE TENSION (cm)
(A) (B) (C)
r
FIGURE 35.—The wetting angle of a film, indicated by the spreading that
occurs, is a measure of wettability.
Several other factors influence adhesion, but their effect has not
been clearly demonstrated for soils. The angle of wetting is prob-
ably affected by the surface roughness of the material. Presumably,
small irregularities act like small capillaries in which water can rise
and thereby increase the film contact. The viscosity of a fluid may
also affect adhesion, since it will influence the movement of water.
Where relatively short loading times are involved, lack of movement
may affect moisture tension or the area of contact in the immediate
neighborhood of the adhering moisture films. The effect of these
factors has yet to be determined for soils.
As was indicated in the discussion of friction, several factors affect
the coefficient of friction defined by equation 19. Obviously, one fac-
tor is adhesion ; furthermore, the effects of adhesion cannot be simply
separated from friction. The problem is usually circumvented by
specifying equation 19 as the apparent coefficient of friction. That
force normal to the sliding surfaces which is due to adhesion is inter-
preted in a change in the coefficient />t. Since experimentally the ap-
parent coefficient rather than the true coefficient of friction is meas-
ured, the apparent coefficient of friction is normally utilized.
Haines ( 169 ) demonstrated the importance of adhesion on the
sliding friction of metals on soil. The relation is expressed in the
form
^' = ^ - tan 8, (29)
500
10 20 30
MOISTURE (PER CENT)
SOIL CYLINDRICAL
SINTERED GLASS CONTAINER
FARADAYS WAX
MOISTURE TENSION
LOADING BUCKET-
0.7 -
1
0.6 - /
0.5 - /
^,x^
0.4
0.3
FRICTION ADHESION LUBRICATION
0.2 PHASE PHASE PHASE
0.1
1 L_i 1
10 15 20
MOISTURE CONTENT (7o)
FIGURE 38.—General phases of soil friction used to identify soil reactions at
different moisture contents. (Nichols, Agr. Engin. (5i6).)
r T
I
r
SOIL YIELDS OR MOVES
DOMINATING FORCES
a
SOIL RESISTS MOVEMENT
TRANSMITTABLE FORCES
Dynamic Behavior Properties Passive Behavior Properties
• Cohesion • Pore Radius
^ Friction • Tortuosity of Pores
# Adhesion • Thermal Characteristics
L
• MIXED FORCE SYSTEM • J
FIGURE 39.—Relation of static and dynamic behavior as a means of character-
izing soil reaction to an applied force system.
are so ill defined and involved that their magnitude and form are
generally not known. Thus, describing behavior in quantitative
terms, even for passive reactions, becomes almost impossible until the
forces can be adequately described.
At any one instant in time, forces such as those from climate can
cause a passive reaction; water entering soil is an example. When
the water content of a soil increases, these same forces can simul-
taneously cause an active reaction such as swelling. Thus, soil con-
SOIL DYNAMICS IN TILLAGE AND TRACTION 59
ditions are often not static with respect to time but do, in fact, change
with time since the forces change with time. The actual condition
of the soil at any one time is, therefore, a function of the past his-
tory of forces. Since this history is generally not know^n, the con-
dition is attributed to such vague unknown quantities as weathering,
microbial activity, crop rotations, and aging. The inability to define
forces and the time dependence of forces are two factors that greatly
complicate describing behavior.
Mechanical forces such as those usually applied to a soil by tillage
tools or traction devices can be more specifically described than can
climatic or biological forces. Furthermore, their action is of rela-
tively short duration and their magnitude can be controlled. Con-
sequently, mechanical forces as a group can be separated and studied
with relative ease. It is this isolated group that is considered in de-
tail in this handbook on soil dynamics. Since mechanical forces that
are universally dominant forces cause soil reactions, behavior proper-
ties rather than static properties describe the reaction. Dominant
forces cause the soil to yield and become active rather than passive
in the soil reaction. A detailed study of behavior resulting from
mechanical forces, therefore, requires a detailed study of dynamic
properties.
To describe behavior in some rigorous manner, both input (forces)
and behavior (the soil's reaction) must be considered. The forces
cause the action, whereas the reaction determines the kind and amount
of change in soil condition. The action and the reaction are equally
important and must be properly described. The inñuence of the
condition of the soil on the reaction must also be indicated. Behavior
properties provide a logical means for describing this influence. A
relation is, therefore, implied to exist between physical and behavior
properties. A unique relation must exist because once a finite amount
of material is isolated for subjection to a force system, its physical
properties along with its behavior properties are fixed and deter-
mined. If these properties change during the action, the change can
theoretically be included in the description of behavior. A unique
though possibly complex relation, therefore, must exist between phy-
sical properties and each behavior property.
The relation between physical properties and behavior properties
can be more clearly seen when two facts are recognized. First,
physical properties primarily identify soil and its condition. They
are always in evidence. Behavior properties are made manifest
only when a specific action occurs. An action is not necessary for
physical properties to exist ; they are fundamental entities in them-
selves. A behavior property is not fundamental in itself ; rather, the
specific action is the fundamental entity and its defines the behavior.
The second fact is that static state and material properties (physical
properties) are independent of each other. A soil may be found
either in a dense state or in a loose, pulverized state. The soil ma-
terial is the same but its state is different. In a similar manner,
different soil materials can exist in the same static state. Therefore,
properties that identify a soil material and properties that identify
its static state are, within reasonable limits, independent of each
other. Furthermore, since physical properties describe the nature
60 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
of a material and its nature determines how the material will behave,
behavior properties must be a function of physical properties. In
other words, in a mathematical equation that describes a relation,
state and material properties are independent variables and behavior
properties are dependent variables.
Researchers have intuitively reasoned that a unique relation must
exist betw^een physical properties and behavior. If such a relation
does not exist, a different behavior would result whenever the same
system of forces is applied in the same manner to a material in a
given condition. Repetitive behavior, however, is observed when
carefully controlled and standardized procedures are followed.
Knowledge that such relations exist has often guided research
efforts in wrong directions. To illustrate, again consider the metallic
wire and Ohm's law. Available knowledge indicates that the resist-
ance of a piece of wire is inñuenced by the length of the wire, cross-
sectional area, chemical composition, and temperature. All these are
physical properties that identify the piece of wire. Imagine the
difficulty of describing behavior when voltage is applied if these
physical properties were to be included in the description without
knowledge of the behavior property, electrical resistance. Once re-
sistance has been identified, however, a logical means for describing
the behavior of current flow exists. First, behavior is described and
a behavior property is identified (Ohm's law and resistance). Then
the behavior property rather than the behavior itself can be studied.
Recognizing the existence of behavior properties and first search-
ing for them is a means of simplifying the problem of describing
behavior. The possibility exists that one physical property or some
simple combination of two or three physical properties might be
directly equivalent to a behavior property. In such instances, the
relation between physical properties and behavior may be found
without first identifying a behavior property. When the complex
behaviors of soil are observed, the possibility of directly relating
physical properties and behavior seems remote. It looul'd thus ap-
pear that the ^ relationship hetween the physical properties that
intuitively must affect dynamic properties and hence soil hehavior
must await means of identifying and quantitatively measuring the
dynamic properties. To fail to recognize this can only lead to con-
fusion and frustration.
3.2 Dynamic Parameters
In the preceding section, dynamic properties were shown to be
capable of characterizing soil as it reacts to applied forces. Inter-
relations betw^een applied forces, dynamic properties, and behavior
are implied in figure 39. The figure also suggests the qualitative na-
ture of these relations. The soil and the applied forces are both of
importance so that once a specific soil is isolated and a specific system
of forces applied, the resulting behavior is determined. The be-
havior must be considered in a restricted form; hence, a statistical
form of result is not produced. The soil and forces thus are primary
and basic, and the behavior is derived from these basic factors.
The forces, how^ever, are the prime movers of any action. They
cause the action, so they may be properly termed inputs to any equa-
tion that expresses the manner of soil behavior. These inputs act
son. DYNAMICS IN TILLAGE AND TRACTION 61
properties of the soil although the exact relation may not be known.
A complete quantitative description of behavior of the soil to forces
can be determined after one has identified, qualitatively defined, and
finally quantitatively assessed dynamic parameters of soil.
Although not indicated at the time, the requirements for a quanti-
tative description of behavior were met when dynamic properties
were discussed in chapter 2. For example, when soil deforms, the
output of the behavior equation is the resulting movement. Strain
was discussed in section 2.3 as a means of describing movement so
that numbers could be assigned. The forces acting in the system
cause the strain ; and stress, which numerically describes the forces, is
the input. Stress-strain equations replace the black box and define
a dynamic property, which reflects behavior. The parameters of the
stress-strain equations numerically assess the dynamic property.
Another example of behavior is shear; the output differs in this
instance, since the output description itself does not need to be
quantized. As indicated in section 2.8.1, shear is a yield condition;
and the output is the failure itself. Although the failure is clearly
described, a number is not necessarily specified. The input for shear
is again the applied forces, and they are the forces acting at incipient
failure. The relation between the stresses at incipient failure and
yield provides the real mathematical function of the black box.
Thus, equation 18 defines the dynamic property shear, whereas co-
hesion and the angle of friction are the dynamic parameters that
numerically represent shear. Both stress-strain equations and shear
yield equations are examples of force inputs that are related to be-
havior outputs by an equation that represents the behavior.
A superficial inspection of the previous analysis may raise the
question of whether forces are always inputs. In shear, for example,
shearing stress may seem to be an"^ output of the behavior since the
shear stress is the maximum value that can be obtained for the situa-
tions under consideration. If the intended purpose of the soil is to
obtain traction, shear stress may be visualized as the limiting value
of traction for the particular situation ; hence, shear stress is of direct
interest and appears to be an output of the behavior. Shear stress,
however, causes the failure; failure does not cause the shear stress.
The confusion is explained by Newton's Third Law of Motion, which
states that every force has an opposite and equal reaction force.
Thus, the shear stress has an opposite reactive stress, and this re-
active stress is the limiting value when the intended use of the soil
is for traction. Behavior and intended use of the soil, therefore,
must be kept separate if the mechanisms in behavior are to be clearly
understood.
Dynamic parameters are not always present in equations defining
dynamic properties. Such a situation arises when the behavior out-
put is not ¡quantized and the force input is simply represented. In
fact, the situation can be so simple that no equation is required.
In tensile failure, for example, the magnitude of the normal stress
completely describes the behavior at failure. Dynamic parameters,
thus, are not basic because they are defined only by the equations that
represent the mathematical models of the behavior. It is the be-
havior and implied dynamic property that are real and basic. For
SOIL DYNAMICS TN TILLAGE AND TRACTION 63
(A) (B)
FIGURE 40.—In situ direct shear apparatus showing shear failure : A, Soft
conditions; B, hard conditions.
addition, strain may not be uniform along the failure surface ; there-
fore, the observed movement of the apparatus does not represent
movement at the failure zone.
Jenkin ( 200 ) demonstrated the importance of distinguishing be-
tween conditions simulating constant volume and those simulating
constant stress. In an extreme condition represented by compacted
sand, he found that the force required to rotate a steel plate in-
creased 270 times when the volume was not permitted to expand
during shear. Confinement caused shear within individual sand
grains rather than between grains, hence the increased force.
Experiences such as Jenkin's indicate that shear should be meas-
ured under controlled stress. Extreme care is required to insure
that the envisioned model is being accurately represented by the ap-
paratus. The precautions and techniques will not be covered here.
The degree to which an apparatus duplicates conditions represented
by the assumed mathematical model depends on the kind of apparatus
and the techniques of its use. As a result, different shear values are
often obtained on the same soil with different apparatus. Various
aspects of their appropriateness will be discussed in conjunction with
individual apparatus.
In all direct shear apparatus, the normal and tangential forces
are measured as well as the relative movement of the apparatus with
respect to the soil. The normal and tangential forces divided by
the appropriate failure area gives the required stresses. Movement
is generally expressed in terms of strain. The data can be repre-
sented as shown in figure 41, where the tangential or shearing stress
is plotted versus the strain for a constant normal stress. The three
characteristic curves represent the range of behavior observed.
Condition A,is characteristic of highly cemented and dried soils or
hard, clean sands that exhibit a high peak strength. Once initial
failure occurs by breaking cemented bonds that reflect high cohe-
sion, the measured stress lessens and the friction component is pri-
marily represented. Condition B represents the other extreme where
the soil is loose and has low cohesive strength. Any number of
intermediate behaviors such as indicated by condition C may be
found.
From data such as indicated in figure 41, the parameters of shear
can be determined. Incipient failure is adjudged to occur either
when the shear stress reaches a definite peak (condition A) or when
the shear reaches a plateau (condition B). A series of shear meas-
SOIL DYNAMICS IN TILLAGE AND TRACTION 67
A
STRAIN
T
B^,.^_
rc
-a
i w
(A) (B)
s - -B-, (30)
where M = torque to cause shear,
r = radius of the plate,
and by assuming that the soil was plastic and thus the stress was
independent of strain.
Fountaine and Payne ( 129 ) developed a portable apparatus to
measure shear strength of field soils. A hollow, circular shear box
was driven into the soil until the top of the box was in contact with
the soil. The soil against the outside of the box was then carefully
excavated before shear failure was measured, so only the soil at the
SOIL DYN^AMICS IN TILLAGE AND TRACTION 69
TORQUE
SHEAR BOX
WEIGHTS
r/m//M//á
bottom of the box was sheared (fig. 43). Markers were placed in
the soil inside the shear box so that they were visible through little
holes in the top of the box (fig. 43). By watching the position of
the markers as the soil was sheared, they concluded that the mass
of soil in the shear box moved as a unit and that the rate of strain
within the sample could be considered uniform.
Cohron ( 79 ) has devised a portable torsion shear of this type
that records the shear data without elaborate instruments.
In an effort to overcome the fact that the outermost elements must
move considerably farther than those in the center, a narrow annulus
has been used as a shear apparatus ( 1^8, 162, 2J^5, 397 ). Shearing
stress is easily calculated for a narrow annulus by using polar coor-
dinates. An elemental area is given by
r dB dr,
and assuming a constant shear stress S acting on the annulus area,
the force on the elemental area is
Srdddr,
The force acts at a distance r from the center so that the moment
at the center of the annulus is
S r^ de dr.
Integrating over the appropriate area gives the total moment, which
has the form
M Sr^ de dr.
J^2 Jo
M - 3
70 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
and solving for S gives
UI
S (31)
277 {n^-r2^y
where S = shearing stress,
M = torque to shear the soil,
ri = outride radius of annulus,
^2 = inside radius of annulus.
Another rotating circular apparatus (fig. 44) has been proposed
2>
in the form of a vane shear ( 7 ). Once driven into the soil, rotation
causes shear of soil along the surface of the cylinder, which is gen-
erated by the vanes. This device may be used at great depths in the
soil without excavations. Measurements may be made at succeeding
depths without extracting the shear device, so that a rather complete
strength profile of natural soil conditions can be obtained. Shear
stress may be computed from the equation
UI
S = (32)
287rr^'
when the vanes have a height-to-radius ratio of 4:1. The vane shear
provides no means of varying normal load, although Evans and
Sherratt ( 113 ) have estimated values of cohesion from vane shear
data.
A more sophisticated shear device has been developed in the form
of a so-called triaxial apparatus (fig. 45). A cylinder of soil is
confined in a thin impermeable membrane, which is tightly sealed
around two parallel circular platens located at the ends of the soil
columns. As long as the membrane is very thin, it has essentially
no influence on the stresses within the sample ( Jiiïl ). When the
entire system is submerged in a fluid and pressurized from an ex-
ternal source, the soil is subjected to a uniform pressure on all sides
that is eq^uivalent to the pressure in the external chamber. Friction
and arching may cause small irregularities in stress distribution at
SOIL DYNAMICS IN TILLAGE AND TRACTION 71
-AXIAL LOAD
AXIAL
CHANGE
PRESSURE
RUBBER
MEMBRANE
DIAMETER
CHANGE
POROUS
PLATE
-H^O
PRESSURE
Lhl^^VOLUME
CHANGE
FIGURE 45.—Triaxial shear apparatus modified to measure diameter strain.
the platens, but there will be considerably less friction and arching
in this system than in any other known system. Following an initial
uniform loading by the hydraulic system, an additional mechanical
axial stress can be applied to the soil by means of a rod that extends
out of the external pressure chamber. By increasing the axial stress
over the lateral stress, shear failure can be caused. Failure occurs
either when a clear fracture is evident or when the diameter of the
sample is increased. At incipient failure the principal stresses are
known, and the results may be plotted in Mohr's diagrams. Values
for C and ^ can be determined from the diagrams, as shown m
figure 24. While the name of the triaxial device indicates that the
stresses can be controlled along three axes, such is not the case be-
cause the two lateral stresses are always the same. Nevertheless, the
triaxial apparatus provides an excellent means for making shear
measurements.
The usefulness of basic shear data depends partly on how it is
incorporated into a soil-machine mechanics. As isolated values, C
and ^ have no practical usefulness other than to characterize soil
conditions. Several attempts have been made to extend the useful-
ness of the basic shear data. Payne ( 329 ) pointed out that much of
the time when soil is being moved, the maximum shear stress only
partly represents the conditions. He reasoned that large strains in
soil would indicate that the magnitude of the shear stress represented
by a plateau, as illustrated by condition A in figure 40, is a more
reasonable estimate of the acting stress. The value of 0 obtained m
72 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
APPLY TORQUE
WAX
PLASTER
CLOD' OF PARIS
FIGURE 46.—A method for determining shear of individual soil clods. (Foun-
taine, Brown, and Payne, Sixth Internatl. Cong. SoU Sei. Soc, 1956 {128).)
S = (36)
(A) (B)
FIGURE 48.—Methods of applying forces to simple beams to cause tension
failure.
the tensile stress at the outer edge of the soil beam at failure. Be-
cause the assumptions are not satisfied even for ductile metals, the
failure stress is designated modulus of rupture.
As shown in figure 48, a beam may be loaded in several ways. The
arrangement at the left provides maximum bending at the center of
the beam whereas that at the right provides a long central test sec-
tion in which bending is uniform. Carnes {61)^ Eichards ( 364, ) ?
and Allison ( 2 ) have utilized the modulus of rupture in the study
of soil crusts. Crusts on the surface of the soil can be collected with-
out undue disturbance. This simple measurement can be used to
evaluate their strength in a relatively undisturbed condition. This
method has been widely accepted and may be considered as stand-
ard {US).
Another method of causing tensile failure has been used by Kirk-
ham, DeBoot, and De Leenheer ( 218 ) for soil core samples. In this
method, a cylindrical sample is compressed laterally until a tensile
failure results, as shown in figure 49. The failure stress of the soil
FIGURE 49.—The indirect tension method for causing tensile failure, (Kirk-
ham, DeBoot, and De Leenheer, Soil Sei. ( 218 ).)
S (37)
TTDÚ
where F — breaking force,
L = length of the sample,
D = diameter of sample,
S = modulus of rupture.
SOIL DYNAMICS IN TILLAGE AND TRACTION 77
The cores must be homogeneous for this method to be reliable. Its
utility lies in the fact that core samples provide relatively undis-
turbed soil samples for evaluation.
Mitchell ( 301 ) evaluated this indirect tension-measurmg method
for samples of concrete, and he concluded that the contact between
the compression plate and the sample was of extreme importance.
As shown in figure 50, different types of shatter result if the contact
CARDBOARD PLATE
50,000
3 40;000
Q 30,000
<
3 20,000
10,000
10 30 50 70 90 10
SOIL BAR
BREAK POINT
CENTRIFUGE HEAD
FIGURE 53.—Apparatus utilizing a centrifuge to determine tensile strength of
soil. (Vomocil, Waldron, and Chancellor, Soil Sei. Soc. Amer. Proc. { 469 ).)
80 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
of the block is confined by the rim of the centrifuge head. While the
sample should normally fail at the center of rotation, it may fail at
some other location and the radius r must be calculated on the center
of mass of that portion of the block. The acceleration force also
moves the water in the soil, and this movement may affect strength
at the center of the block. The tensile strength of soils may be very
high—indeed, as high as that of some building materials—but this
strength disappears very readily on wetting. Moisture movement,
therefore, can be important.
In general, tensile strength determinations have been made with
the aim of evaluating the maximum tensile stress of the soil. Hen-
drick and Vanden Berg ( 181 ) have also measured strain in the
failure area so that tensile stress-strain relations could be obtained.
A sensitive linear variable differential transformer was used to
measure very small strains before soil failure. As a result, strain
energy could be determined from an integration of the area under
the experimental stress-strain curve. Strain has also been measured
in conjunction with the beam and the indirect methods of measuring
tension. Whether this can be done with any degree of ease in the
centrifuge technique has yet to be determined.
One of the limitations of tensile strength expressions has been
the basis on which to express the force. Since only the solid or solid-
and-liquid phase can support stress, the force should be expressed
on this area rather than on the cross-sectional area of the soil sample.
The failure area is known when the soil is in either a completely solid
or a saturated condition but is not known when the soil contains air-
filled pores. Eeporting force on the basis of the void ratio is some
improvement but is still not exact. Unfortunately, no method has
yet been developed for determining the area that transmits stress.
A suitable method must be able to distinguish the bonds that have
continuous materials from pseudobonds that have only contiguous
materials. Low energy vibrations may serve as a means by which
to evaluate, without destroying, effective bonding areas in soil.
Recently, Vomocil and Waldron ( 1^68 ) have attempted to compute
the tensile strength for an unsaturated system of glass beads and
water based on the area of the moisture films and the soil moisture
suction. Following the approach of Fisher {119, 120), the pro-
jected wetted areas of spheres were used as a basis on which to
compute the tensile strength of a mass of wet beads. By using the
geometry of the system at different degrees of moisture saturation, it
was possible to determine a series of areas over which the tensile
force must operate. When these areas were multiplied by the
equivalent soil-moisture suction as predicted from capillary theory,
the computed tensile strength of the mass of spheres agreed with
the value determined from measurements. Unfortunately, computa-
tions based on the measured soil-moisture suction predicted tensile
strengths as much as 40 times greater than those measured. Thus,
there seems to be no simple way to relate soil-moisture suction to
tensile strength, even in a simple system.
3.2.7.3 Compression
As discussed in section 2.8.2, compression is a failure condition
associated with a change in volume of soil. Eecall that failure was
SOIL DYNAMICS IN TILLAGE AND TRACTION 81
10 15 20 25 30
SURFACE PRESSURE (psi)
P' expresses the equivalent load on the soil during shear at different
water contents. Its value depends on the assumed linear relation
described earlier.
SOIL DYlN'AMICS IN TILLAGE AND TRACTION 85
Greacen ( 161 ) lias proposed that the strength of the soil be
described in terms of the maximum pressure that the soil can with-
stand without furtlier consolidation. He proposed that this load
bo expressed logarithmically and defined as the equivalent strength
of the soil log S. In saturated clays, where the angle of friction
can be considered to be zero, the shear strength in a continuously
straining soil may be assumed to be due to the applied suction P\
By definition, log S — log P' in equation 45 ; hence, the possibility
exists of calculating the equivalent strength of the soil from moisture
data. Plots of equivalent soil strength, as estimated from log ^ =
log P', versus the logarithm of the measured moisture suction during
shear did not produce a unit slope; lience complete agreement was
not found in the proposed relation. That the results did not agree
might be explained in part on the assumption that mechanical and
suction forces are equivalent. Kesults of McMurdie and Day ( 282 )
and Vomocil and Waldron (468)^ which Avere discussed earlier,
indicate that this assumption may not be valid in unsaturated con-
ditions.
Another measurement that appears to be particularly adapted for
evaluating plastic strength of soft, wet soils is the Jourgenson
squeeze test ( 487 ). The apparatus used resembles a shear box with
open ends (fig. 55). The soil is placed in the box and squeezed by
FIGURE 55.—Apparatus used for the Jourgensen squeeze test. Open ends per-
mit the soil to flow out under pressure. (Waterways Experiment Station
(^87).)
a vertical stress that causes the soil to be extruded from the open
ends of the box. The shear strength of the soil is calculated by
means of the equation
Pa
S = (46)
BU (l + ira/L)
where P normal load on upper plate,
B width of sample, •
a % sample thickness,
L % sample length.
The strength of soil determined by this method is not significantly
different from that determined by other methods. In thé Jourgensen
test, the magnitude of stress is the plastic parameter.
86 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
A pure shear device has been used by Murayama and Hata ( 306 )
to study the weakening of soil by repeated manipulations. The de-
vice was a shear box with movable walls that permitted the box to be
tilted back and forth. The two horizontal surfaces were kept a fixed
distance apart so that the volume of soil in the box did not change
during the tilting (fig. 56). Aö* shown in figure 57, weakening in-
(A) (B)
FIGURE 56.—The nature of shear in the pure shear apparatus where the
vertical angle is the tilting angle.
2 3 5 7 10 20 30 50 100 200
NUMBER OF REPETITIONS N
FIGURE 57.—Effect of the number of shearing cycles on the degree of weaken-
ing of soil (Murayama and Hata, 4th Internatl. Conf. Soil Mech. and Found.
Engin. Proc, Butterworths, London {S06).)
tion within the mass of soil. As discussed in section 2.9, ri<iid body
soil movement is often observed. Failure conditions are again of
interest, and mathematical models are envisioned to describe rela-
tions between forces of interest. Friction is one dynamic projjerty
involved in rif^id body soil movement. Friction appears as soil-soil
friction wlien two rifiid bodies of soil are involved, and it appears as
soil-material friction when one soil body is replaced by some other
material such as a steel tillage tool. The relative importance of the
two frictions in any soil-machine relation depends on the reaction,
but generally both must be considered.
The difficulty in measuring soil-soil friction as defined in equation
1!) is to distinguish between the soil-soil friction angle i// and the
shear failure friction angle <^that is, the slope of the line in the
mathematical model. That the two angles refer to t\yo distinct
))henomena was discussed in section 2.!). On the basis of these dif-
ferences, t|/ is usually measured by drawing a rigid body of soil having
a known cross-sectional area over a soil surface in a similar condi-
tion. The methods for measuring shear failure thus lend themselves
to soil friction measurements. Both direct apparatus and torsion
apparatus are used, and generally the tangential stress and strain are
simultaneously measured at various normal loads. The results are
expressed as shown in figure 41, and the numerical values for equa-
tion 1» can be determined after the shearing stress reaches a plateau
where clear separation on the failure surface is assumed to occur.
The limitations and errors of the measurements are the same as dis-
cussed in section 3.2.1.1. Thus, soil-soil sliding friction is represented
by the single dynamic parameter fi or tan ip.
Soil-material (soil-metal) friction can be identified as the angle of
sliding friction ô or the tangent of the angle /u,', which is determined
FIGURE 58—A simple slider system used to determine the coefficient of slidinK
friction.
88 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
-INSTRUMENTED BEAM
TINE
-SENSITIVE
PLATE
FIGURE 59.—The NIAE Friction Apparatus. The view on the right is a plan
view showing how the sensing area was subjected only to forces caused by
the soil sliding. (Rogers and Tanner ( S12 ).)
value of
Critical
^^
o
5
Çjj TtH CO Tt^ 00
CO
eg
value of
Critical
II
c¿ CO CO o ïO
û^' T-í rH TH TtH
value of
Critical
CO
o
o Soococo
5
ÍSjíO COTí^CO
CD
en
is« value of
VA
A
«H
Critical
•'^* o iO o o
0 ^
C CO 00 l>; C5 W C5
û^" TH TH -A TjH
4.J
â>»
-M
A
% CÖ
OJ
u
value of
Critical
0 >H
«H p
MíO
value of
Critical
■gil •^' KÎ lo o o
do Oi TH 00 Oi
(^' r-i <N* TH Tti
•^ CO
'?
Cü -M
n^
-M
S
'5* ^^ CO
O» ^•OrH
.iH
iD
•
o; ,r|
pa <M **H
©0
>.^
00 t>5
u fl ,^^
CÖ
V^ Tt^ O CO O CÖ a OJ
a 0a .S
íO Ç55CO CO CO (M
0 O -M
0
03
a
03
P
-4J 02 ^=4
Co yj i3
«5
■e-
1 i3
a a '00
-M 03
05 CO CO CO CO
+j ci
03 A
A •^ 0)
a 4->
Ö
P>»
2 «O 0 ^
^
«H «H
0 0
0 o; Prl
o o Oí! (M C<1 C^ '^
C3 f3 C5
m
^^ 0
H ç^ m
SOIL DYNAMICS IN TILLAGE AND TRACTION 93
The attractive force of adhesion is determined by two factors:
first, the actual strength of attraction of a unit area of bonding;
second, the actual area of the attractive bonds. The pressure applied
to the metal surface inñuences both of these quantities so that
evaluation of the results is difficult. The adhesive values measured
by the metal plate technique must have an initial conditioning load-
ing which duplicates the actual loading of the system under study
or the results will have little meaning. As shown in table 9, the
20 40 60 80 100
STATIC PENETROMETER RESISTANCE (Kg/cm*)
Oi
SOIL NOT
Wk PERMITTED TO DRY
SOIL PERMITTED
TO DRY
FORCE (lbs)
JP
A 10:
b
FiGUBE 62.—The relation between applied load and ratio of settlement to
diameter for four plate diameters. (Amer. Soc. for Testing Materials {9 ).)
LOG z
T
Í
FIGURE 64.—Graphical solution of equation 52 to give values of kc and k..
either increase or decrease, and the change is often much larger than
in metals. An example of the phenomenon is the manipulation of
snow into a snowball, which results in a severalfold increase in
strength.
In all materials, induced strength is affected by the physical
properties of the material as well as by the applied forces so that
both must be considered. Stress-strain relations provide one suitable
means for mathematically undertaking a simultaneous consideration
of both factors. As has already been stated, however, accurate
stress-strain relations are not available for soil. Attempts to measure
induced strength by other means have not successfully identified
parameters that assess the dynamic properties of induced strength.
Eather, the attempts have assessed the interaction of the manner of
application of the applied forces and the resulting change in strength.
Consequently, the results are composite assessments of the behavior.
Shear has been used in nearly every attempt to assess changing
strength. The parameters of equation 18 (discussed in sec. 3.2.1.1)
thus become the measures of induced strength. One exception has
been made in connection with traíRcability, in which soil resistance is
measured with a cone penetrometer before and after the soil is
subjected to a specified amount of manipulation that expressly rep-
resents vehicular traffic on the soil. The second measurement assesses
any change in strength—that is, induced strength. In shear, the
change of strength is usually represented as a change in cohesion,
since the angle of internal friction generally is assumed to remain
constant. T i
Shear may be measured both before and after some standardized
manipulation of the soil to assess the change. Since shear is actually
a manipulation of soil, a special interpretation may be applied to
the shear data to determine two values of cohesion. An example is
Payne's residual cohesion (sec. 3.2.1.1). Two values of cohesion
were developed from his initial measurements without any special
manipulation of the soil. The second value, of course, was based on
a decrease in strength because of a disruption of bonds in the soil
during shear.
A clear example of the before and after measurements is available
from Transportation Research Command data ( Í32, J^SB ). Two
distinct values of cohesion were proposed to reflect induced strength.
Precollapsed cohesion would reflect the initial strength, whereas post-
collapsed cohesion would reflect the strength after a load had been
applied. Any difference between the two values would be a measure
of induced strength. The only difficulty with the concept lies in
measuring postcollapsed cohesion, which obviously depends on the
load that is applied. Since the postcollapsed value can be deter-
mined only after the applied load has been removed, a misleading
value of cohesion can be obtained. Thus, from a practical standpoint,
the concept has a limitation because generally there is little need to
know the strength after the load—say, a vehicle—has passed over the
soil. Furthermore, the possible interaction between the soil and the
apparatus that measures shear limits the practical aspects of the
To avoid these difficulties, an average or relative cohesion Creh
concept.
102 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
was proposed that would reflect the desired cohesion in the appro-
priately loaded state. An example of the concept is shown in figure
65, where sinkage of various plates into snow is correlated by assign-
10 - 10
Orel 12 in DIA. PLATE
8 l2mDIA.PLATEv
9¡nDIA. PLATE \ V
r ■ 9 in DIA.
cr|ü 8 - -o-6 in DIA.
1 . -x-3in DIA.
PLATE
PLATE
PLATE
/f
ü
6 6 in DIA. PLÀTEv X y 6
-^^T^^ 1 . 1 . 1 .
0 0 ■^^i 1 • 1 . 1 ,
FIGURE 65.—A, Pressure on plates versus sinkage in deep dry snow. B, Same
data as in A corrected for relative cohesion Crei. (Transportation Research
Command ( ^55 ).)
<
I
Z
<
y = A ^(^:H^ (55)
- \,^MODULUS OF RUPTURE
20 - ^X y^ -
10 \ y/
: \, / I
UJ \ ^COEFFICIENT - in
<
Q:
ID \ / OF ABRASION.
2 - CQ
<
1 _ /\ LL
a: " / ^ ' O
Ü-
o
- ^ \
.2 / .02 ^
\
.1 - / \ : .01
o / \ O
/ \ ■
o
.02 . . . \ ...002
.001 .002 .005 .01 .02 .05
E
3 4000
O
"^ 3000
z
g
<
g 2000
<
1000
1500 2500
ABRADER (gm)
fTTTTTTnir^
1 ^
Í 3
UJ
I 2
DIAMETER (Microns)
•
10
--
•
- •
3 ® --^ •
z •
o •
01 •
H
UJ r^
S 4-
Û- •
/
2 - /
./ 1 • 1 L.. 1
10 20 30 40 50 60 70
KINETIC ENERGY (Ft Lbs)
.13
.11
if)
O .09
.07
<
NORMAL LOAD .05
12 Lbs.
.03
.01
2 4 3 9 15 21 27
TIME OF WEAR (min) NORMAL LOAD (lbs)
BEFORE RAIN
0.5 AFTER RAIN
0.3
3.3.4 Movement
The condition of the soil at one time as compared to another
might be described in terms of its overall location or configuration.
This is not the same type of description that might be used to de-
scribe the internal arrangement or soil structure—that is, the
size and shape of the aggregates and how they are packed together.
Instead, the relative location of different parts of the general soil
mass are of interest. As an example, a lowering of the surface of
the soil would indicate that subsidence or compaction of some form
had occurred. In other cases, the soil may be inverted or moved
as in plowing, harrowing, or land forming and it would be import-
ant to trace the movement. This movement can be seen in a plowed
furrow, yet the movement is not easy to describe.
A study of the movement in detail requires the use of some ref-
erence system so that the initial and final locations may be identified.
Strain, as discussed in section 2.3, provides a rigorous model for
describing movement and final location in terms of initial locations.
As was pointed out, both rigid body movement—including rotation
and translocation—and strain are required for a suitable description.
But the mathematics becomes very complex even when simplifying
assumptions are made, and little progress has been made in using
strain to describe movement. Kather, initial and final locations are
identified in some suitable reference system so that gross movement
is observed. The movement is usually not even directly associated
with the soil itself as a means for characterizing the soil—that is,
a movement dynamic property. Instead, the movement is used to
qualitatively observe the behavior of the interaction of soil and
some device or machine.
The initial and final locations of soil have been assessed in several
ways. Nichols and Eandolph ( 320 ) placed aluminum foil marker
strips in the soil so that movement of the soil would move the strips.
The degree to which this material would alter the soil reaction would
depend on the strength of the aluminum and the amount of strain
in the soil. Other workers have used noncontinuous tracers such as
colored beads, brass rods, sticks, gypsum, and coal dust {90, 138.
200, 267, 418 ). Both line and grid patterns of these materials have
been used to mark the soil, depending on the object of the study.
These tracers must be placed in the soil very accurately before soil
movement is initiated. By carefully cutting the soil mass apart
after a specific operation has been completed ( 100, 251, 300 ), the
final position of the markers can be determined. The movement
of the soil can be inferred from a simple plot of the tracers. Figure
74 shows the nature of soil movement caused by a moldboard plow.
The description does not indicate the paths traveled by the in-
dividual markers.
Figure 75 shows a glass-sided box technique in which the soil
particles themselves serve as tracers. Although this technique has
limited application, it provides a means by which a continuous path
of movement can be observed. The visible lines are caused by the
movement of individual particles or aggregates of soil along the glass
side of the box. The particles scrape the fine layer of aluminum
114 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
I
^ 0 L
"
UJ 3 - >
o
<
-1
2 - LU
t-
13
• 4
1 - z
X
(B)
3
DISTANCE (Ft)
from the glass so that the path of the particle in contact with the
glass is recorded.
Figure 75 represents the total movement that occurred in the
soil as a rectangular probe was forced into the soil along the glass.
Movements do not occur simultaneously and only a series of such
pictures would show the progress. The extent to which this move-
ment would occur in the soil if the glass side were not present has
been questioned. When the glass is lubricated, possibly the move-
ment of soil near the glass is relatively similar to that which would
occur if the glass were not present. When glass is not used or when
the tracers are not visible, their location must be determined by
other means.
The use of lead shot in soil has permitted determining their pro-
gressive location by means of X-rays without destroying the soil mass
( 217^ iOJf ). Photographs can be prepared to show the location of
the individual shot after various increments of soil movement.
Swietochowski, Bors, and Przestalski ( 1^17 ) used radioactive shot
from which gamma rays emanated. These rays impinged on photo-
graphic paper to mark the location of the shot sources. Unless the
location of each shot is determined as movement progresses, their
direction of movement cannot be determined. Fairbanks ( IH )
used plowshares containing radioactive materials so that particles
which were worn from the plow could be detected in the soil with
radiation detectors. Because the worn material would be attached
only to soil that slid on the plow surface, soil from that particular
layer could be identified.
Phosphors are materials that emit visible light during or after
radiation by ultraviolet light. Staniland ( Í09 ) and Wooten, Mc-
Whorter, and Eanney ( 512 ) have used ñuorescent tracers to study
the efficiency of cultivation methods. Under field conditions, an
ultraviolet light has been used at night to illuminate ñuorescent
tracer materials so as to be able to study the degree of soil mixing
caused by different tillage tools.
Martini ( 293 ) used plaster markers to identify the initial center
of gravity of a section of a furrow slice. He caught the plowed
soil in a large pan placed in the bottom of the furrow and then
balanced the pan to redetermine the center of gravity of the final
position of the soil mass from the test section. He considered the
difference in the location of the center of gravity as an index of
the soil reaction to the plow.
The configuration of the soil must often be altered into some
specific form for a special purpose. A description of the surface
is easy when the shape is simple. The most accurate method used
is to measure the elevation of each point of the surface in reference
to some base point. A surface contour map can be constructed from
the data. The field survey method is the most practical for describ-
ing large areas such as those in which land forming, grades, and
water courses are of interest. Mech and Free ( 296 ) and Soehne
( Jt02 ) have studied relatively rough plowed surfaces by a version
of this method. At least one device is commercially available for
measuring the irregularities in surface profiles of highways and
airfields ( ^7 ), but this apparatus is not adequate for most detailed
116 AGRICULTURE HANDBOOK 316, U.S, DEPT. OF AGRICULTURE
4.1 Introduction
Tillage tools are mechanical devices that are used to apply forces
to the soil to cause some desired effect such as pulverization, cuttmg,
inversion, or movement of the soil. Tillage tools usually produce
several effects simultaneously. The ultimate aim of tillage is to
manipulate a soil from a known condition into a different desired
condition by mechanical means.
The objective of a mechanics of tillage tools is to provide a
method for describing the application of forces to the soil and for
describing the soil's reaction to the forces. An accurate mechanics
would provide a method by which the effects could be predicted and
controlled by the design of a tillage tool or by the use of a sequence
of tillage tools. Furthermore, the efficiency and economy of the
tillage operation could be evaluated from the mechanics. A tho-
rough knowledge of the basic forces and reactions is required to
develop the mechanics. Such knowledge is not available at present,
and soil reactions cannot even be predicted, let alone controlled. As
a result, an operation is performed, the conditions are arbitrarily
evaluated, and additional operations are performed in sequence until
the conditions are adjudged to be adequate. Thus, today, tillage
is more an art than a science.
Progress has been made, however, in developing mechanics where
simple tools or simple actions are involved and where forces and re-
actions can be described. This chapter presents several approaches
that have been used to develop simple forms of soil-tillage tool
mechanics. Only homogeneous soil conditions are considered. Al-
though this approach is completely unrealistic, it does not negate the
results of the studies. Complete knowledge of reactions for a homo-
geneous soil will provide a basis for solving problems dealing with
layered soils. Interactions of importance will probably occur, but
they should not present unsurmountable obstacles. The approaches
discussed in this chapter do not represent any final solution of the
problems that are posed. The approaches, however, do represent
those that have been utilized and those that may contribute to the
development of a successful mechanics of tillage tools.
4.2 The Reaction of Soil to Tillage Tools
The reaction of soil to a tillage tool can be quantitatively described
only by a mechanics. Visualize the soil as a continuous semi-infini-
tive mass composed of air, water, and solids arranged in some homo-
geneous manner. As a tool advances in the soil, the soil reacts
to the tool and some action occurs. For example, the soil may move
as a mass, tlie solids may displace the air or water, or the solids may
117
118 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
and bulk modulus and shear modulus are behavior parameters de-
fined by the equations that characterize the material. Simultaneous
consideration of the three behavior equations (note that stress-strain
behavior is simple even though the equations are mathematically
complex), together with all of the associated boundary conditions,
provides a basis for a mechanics. The solution of the system of
equations will account for the possible interaction. Interactions
can, therefore, influence the procedure for developing a mechanics.
Certain generalizations for developing a mechanics based on be-
havior equations can be concluded from the discussion of the New-
tonian examples and can be summarized under three points: (1)
The action to be quantitatively described must be defined. (2) The
behavior involved in the action to be described must be recognized.
(3) In most circumstances the behavior must be incorporated into a
mechanics that describes the action. Point 2—recognizing the be-
havior involved—is by far the most difficult of the three points,
because recognition usually implies selection. Defining the action,
however, must be accomplished first, so it is discussed first.
The action to be described is defined by interest from outside the
action. A problem to be solved, curiosity, or merely a quest for
knowledge are sources of interest. In the example of the projectile,
interest determines whether the path of motion of each mathematical
point of the projectile must be described or whether only the path
of motion of the center of mass must be described. No set procedure
can be established for defining an action because the procedure
usually embodies simply defining the problem. Personal interest
and the nature of the action itself will influence thé definition.
Until the action (defined here as the doing of something) has been
at least qualitatively defined, however, the problem of quantitatively
describing the action cannot be undertaken.
In any action known to man today, more than one behavior is
involved. Behavior is defined here as any phenomenon that can be
identified, isolated, and studied so that a behavior equation can be
written to quantitatively describe the phenomenon. Thus, Ohm's
Law, stress-strain equations, and Newton's Second Law of Motion
are examples of behavior in the sense defined here. We know from
available knowledge that a rigid body is not really rigid. Strain
always occurs so that the concept of rigidity is relative. Similarly,
a body that strains and is assumed to be continuous is really not
continuous. The body is built up of crystals or aggregates formed
from molecules that are, in turn, formed from atoms. When the
action to be described concerns so-called rigid body movements, all
of the smaller behaviors (smaller because of the physical size of
their sphere of influence) can usually be ignored. Even when the
action to be described is in the realm of the atomic dimensions, today
we know that the atom has a nucleus and the nucleus itself is being
subdivided. Presumably, behavior equations must exist for particles
within the nucleus; therefore, any action known today probably
involves more than one behavior.
No unique system or structure of behavior equations exists. Such
a structure can be developed only when matter itself can be abso-
lutely defined. If, for example, the makeup of the nucleus itself
were precisely determined, matter could perhaps be absolutely de-
SOIL DYNAMICS IN TILLAGE AND TRACTION 121
SOIL CONDITIONS
AND FORCES
Soil Behaviors
f:UJmmm.^mM Separated
Repetitive
Soil Behavior Development of
Phenomenon
Recognized Individual Soil Behavior Mechanics
Observed
Equations Developed
Soil Behavior
Equations Combined
NEW
SOIL CONDITIONS
l//fy7//r^:///^-///^////^/r
thetical forces that act on the segment of soil as it reacts to the ad-
vancing tool. In essence, the behavior outputs have been specified.
Locating failure surfaces specifies the orientation and location of the
force inputs that cause the assumed behavior. The magnitude of the
forces has not been specified. Forces CFi and fJiNi are due to soil
shear and are those present at the instant incipient shear failure
occurs. Forces due to soil-metal friction {iJ/No) and acceleration
{B) are also present. Soehne visualized a pure resistance of the soil
to being split by the cutting edge of the tool {kh). Thus, in prin-
ciple, the simple behavior outputs have been specified in figure 78,
and they represent the complex reaction of the soil to an inclined-
plane tillage tool.
By using the notation in figure 78, an equilibrium equation can be
written for the forces in the horizontal direction acting on the in-
clined tool. The forces on the tool itself are not shown in figure 78,
but they would be the forces reacting to those acting between the
soil and tool and the draft. Equilibrium gives
TT = TV^o sin 8 + ix'No cos Ô 4- hh, (57)
where W — draft force,
^' = coeiRcient of soil-metal friction,
No — normal load on the inclined tool,
k — pure cutting resistance of soil per unit width,
h — width of tool,
8 = lift angle of the tool.
Soehne reasoned that the pure cutting resistance of the soil is small
and becomes important only when stones or roots are present or the
cutting edge of the tool is dull. In the absence of such situations,
the cutting component of the total force might be considered neg-
ligible so that a specific resistance of the soil W* can be defined and
may be indicated as
128 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
TF* = W-kb,
TF* is the tool resistance without the cutting component
TF* = No sin 8-i-NofJi'cos d, (58)
Concentrating now on the soil segment rather than on the tool,
the vertical forces can be summed and placed in equilibrium. With
the notation and the relations shown in figure 78, equilibrium gives
G-No (cos d- fji' sin 8) - ;V^i (cos ß- a sin ß)
+ (OF^ + B) sin ß - 0, ^ ^ ^ ^^ ^^^^
where G = weight of the soil segment,
Ni = normal load on the forward failure surface,
ß = angle of forward failure surface,
fi = coefficient of internal soil friction,
fi' = coefficient of soil-metal friction,
Fi = area of forward shear failure surface,
O = cohesion of soil,
B = acceleration force of the soil,
8 = lift angle of the tool.
The horizontal forces on the soil segment can be summed and placed
m equilibrium from the relations shown in figure 78 to give
No (sin8 + /A^cos8) - Ni (sinyS +/.c cos^) - (OF^ + B) cos^ = 0.
(60)
Equations 58, 59, and 60 can be used to eliminate the normal forces
No and N^. No can be found from equation 58 and substituted into
equation 60. Kearranging terms and solving for Ni gives
¡V, = W-~ (OF,+ B) cos ß
sin y8 +/x cos )ß • ^^^)
Substituting for No and Ni in equation 59 gives
^ TTTT* COS b — uf sin 81
rcos^-usinßl
. ^^ ^ + {CF^ + B) sinyS = 0.
Lsin)ß + /xcos/3j \ ^ / H
Expanding and rearranging terms gives
^y^ pQS 8 - //.^ sin 8 ^ cos /3 - /x sin ^Sl ^ OF^-hB
[sin 8 + ^'cos 8 sin)8 + /xcos)8j ~ sin ß-h fjL cos ß'
and by letting
- ["cQS 8 - ^' sin 8 cosß- fji sin ßl
[sin 8 + /x' cos 8 sin yS + /¿ cos^8 J'
sm jS
cos (6 + ß)
Li = d
sin^S
L2 = 6?* tan 8.
130 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
"^ = itß- m
The acceleration force B is the only item in equation 62 that re-
mains to be specified. Newton's Second Law of Motion provides the
behavior equation, and it can be written
5 = m^, (65)
m = -^hdtvo, (66)
28r
24 24
d = 20 cm.
• MEASURED VALUES
'S 20
^ 5 20
ILÜ
ÜJ Ü
Ü
16 Z 16
z
<
h-
2
¡CO
/d = l5cm.
(0
V)
1? ijj 12
UJ ir
Q:
_J; 8 o
O
en
4-
-j -L.
5 10 15 20 10 20 30 40
FURROW DEPTH (cm) CUTTING ANGLE (»)
(A) (B)
FIGURE 80.—Soil resistances (draft) are measured and calculated for an in-
clined tillage tool operating in a sandy soil. ( Soehne, Grundlagen der Land-
technik (398).)
FIGURE 81.—Free body diagram showing sliding force components due to fric-
tion and adhesion. (Rowe and Barnes, Amer. Soc. Agr. Engin. Trans.
( 376 ).)
SAND CALCULATED
MEASURED
E
Û-
FiGURE 83.—The effect of depth of cut d and lift angle ô on the draft force
of an inclined tool. (Kawamura, Soc. Agr. Mach. Jour. (Japan) ( 2ÖP ).)
84 also shows a clever means that was used to determine the average
shear angle experimentally.
Kawamura noted that as the tool approached the location of the
Ames dial, the surface of the soil rose linearly (slope ^ 0.001-0.01)
with the advancing tool until a critical range was reached. During
the critical range the surface rose at an increasing rate as the tool
continued to advance (relation was a curve). After the critical
transition range had been passed, the relation between the rising
surface of the soil and the advancing tool was again linear but at a
much higher rate (slope ^/ 1.5) than before the critical range.
Kawamura believed that the last linear movement was due to the
rigid block of soil rising on the inclined tool, whereas the initial
linear movement was due to the elastic behavior of the soil. The
critical range, however, coincided with the formation of the shear
surface and reñected the transition state between the time when the
block of soil was part of the soil mass and when it was completely
separated from the soil mass. The shear surface could thus be ap-
proximated by a straight line extending from the tip of the tool to
the soil surface where the critical range began. The average angle
of the shear surface could be determined from the relation
tan ß = -y-,
with values obtained from figure 84.
In other aspects of the work, the lift angle of the tool was ob-
served to influence the mean shear angle so that the shear angle was
not a constant for a given soil condition. Figure 85 shows the
measured relation for two depths of tool operation. The data indi-
136 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
40
_ 30
20
10
10 20 30 40 50
8 n
FIGURE 85.—The relation between the lift angle (5) and the average shear
surface angle (ß) for soil reacting to an inclined tool. (Kawamura, Soc.
Agr. Mach. Jour. (Japan) ( 209 ).)
cate that the average shear angle is influenced by both the depth of
tool operation and the lift angle of the tool.
Kawamura used two theories from the so-called classical soil
mechanics in attempting to calculate the draft force. Both theories
are based on one behavior equation—yield by shear as represented
by equation 18. In the Kankine theory, the shear angle is a constant
given by equation 70. By considering only the mass of the soil and
the shear failure stress (ignoring the magnitude of the normal
stress), the weight of soil and total shear stress on the failure surface
can be calculated. Equilibrium conditions then permit determining
the draft component. In the Coulomb theory, the same procedure is
followed except the shear angle is given by a complicated relation
involving the lift angle and the soil friction angle. Neither theory,
however, was sufficiently accurate to be acceptable.
Since the observed shear surface was a curve and the angle of the
shear surface varied with the tool lift angle, Kawamura ( 210 ) at-
tempted to use plastic equilibrium to calculate the observed phenome-
non. He reasoned that the inclined tool often created a stress state
in the soil different from that required for applying either the
Coulomb or the Eankine theories. He envisioned a transient region
between the tool and the conditions that were accurately represented
by the Coulomb or Eankine theories. At large lift angles, the
Coulomb theory predicted the measured draft closely but the Eankine
theory predicted the observed shear angle better than did the
Coulomb theory. Thus, when small lift angles were used, the devia-
tions between measured and calculated draft could be attributed to
the transient stress region. If the transient region could be con-
sidered to be in a state of plastic equilibrium, the condition could
be properly described by using available plastic flow theories.
The mathematical details for using plastic flow theory are too
long and too complex to be presented here. A discussion of the
principles, however, will illustrate the procedure Kawamura used
to develop a mechanics. When all displacement lies in a plane in a
stress-strain situation, state of plane strain is said to exist. Such a
situation occurs for a wide inclined tool, since all displacement can
SOIL DYNAMICS IN TILLAGE AND TRACTION 137
/m'^7/m//r7wwmmmr-
FiGURE 86.—Polar coordinates used to represent stresses on a slip line caused
by an inclined tool, ß' is the angle between the radial direction and the
?'^fn^\ ^^^ ®^^^ ^^°^- (I^awamura, Soc. Agr. Mach. Jour. (Japan)
24 r
22
CALCULATED-
20
^
< 16
Û
14
12-
10
10 20 30 40 50
S {")
FIGURE 87.—Calculated and experimental draft values for an inclined tiUage
tool. (Kawamura, Soc. Agr. Mach. Jour. (Japan) (210).)
and calculated draft values over a range of lift angles. The agree-
ment is reasonably good, but more data are required to verify the
procedures. Eecall that Kawamura has not considered soil-metal
friction or acceleration forces, and their consideration might improve
the accuracy of the mechanics. On the other hand, he has recog-
nized and considered the curved failure surface and the possibility
of a varying stress distribution on the failure surface. Kawamura
concluded that his procedures represented and explained observed
facts better than any theories that were available at that time.
While his procedures are mathematically complex (primarily be-
cause of permitting a nonconstant stress distribution), they do repre-
sent a mechanics for inclined tools.
SOIL DYNAMICS IN TILLAGE AND TRACTION 139
(A) (B)
FIGURE 88.—The nature of soil failure caused by a vertical tool in a firm soil :
A, Side view; B, plan view. (Payne, Jour. Agr. Engin. Res. ( S29 ).)
For wide tools, the side effect can reasonably be ignored since their
area is small compared to the bottom failure surface (fig. 88). A
tool was considered to be a wide tool when its width was at least
twice its depth of operation. The classical Rankine theory can be
modified to represent the curved failure surface shown in figure 88, A,
When friction is present between the soil and the tool, the directions
of principal stresses do not remain horizontal and vertical. At the
surface of the soil, the principal stresses must be horizontal and
vertical; but their orientation rotates as one proceeds downward
along the failure surface to the vertical tool. The rotation results in
the curved surface. The shape of the curve has been demonstrated
to be a logarithmic spiral, and methods are available to determine
the stresses and the actual arc of the logarithmic spiral ( 4^7 ).
Payne concluded that the modified Rankine theory represented wide
tools with reasonable accuracy.
When narrow tools are considered, side effects can no longer be
ignored. Furthermore, Payne reasoned that shear failure surfaces
must exist which pass along the sides of the tool as well as the bottom
of the tool. Such surfaces will interrupt the bottom curved failure
surface shown in figure 88, -a, and these surfaces will be at Í ~7" — "^ )
to the principal stress. In narrow tools, Payne further reasoned
that the vertical shear surfaces would intersect each other as well
as the bottom curved surface. Thus, a wedge of soil immediately
adjacent to the tool would be isolated from the rest of the soil block
that was sheared from the soil mass. In wide tools, the vertical
shear surfaces would not intersect each other so that the wedge
would never be formed. The wedge changes the boundary condi-
tions for the Rankine theory so that the failure zone of soil can be
140 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
WING
WEDGE
WING
(B)
FIGURE 89.—The nature of soil failure for soil reacting to a narrow vertical
tillage tool. (Payne, Jour. Agr. Engin. Res. (52P).)
DA andW
(A) (B)
HORIZONTAL VERTICAL
DIRECTION OF TRAVEL
. ^
DR Sin 8
1 t. Sin^M
SRCOSÖ Sin(a+X)
1
Sc Cos ß Cos X
DRCOS 8 SpSinö ScSm)3
< T COSÖy
Se Cos^ CosX BcSinr
S-Cosö Sln(a + X)
B^ COS(<^ + T)
(A) (B)
HORIZONTAL VERTICAL
W + DR sin S-^DA-^
A + Be sin r = Tsin 0^.
BcSiiiT
+ BR COS (<^
(é + T) + ^ [/S'A sin 6 + Sc sin )8]. (73)
From equilibrium of forces on the tool (not shown), DR COS 8 will
be the draft. One unknown can be eliminated in a simultaneous
solution of equations 72 and 73. The angles 8 and <^ are dynamic
soil parameters that can be measured, but the remaining unknowns
must be determined from other relations. Equations 72 and 73, how-
ever, are a basis for developing a mechanics of narrow vertical tools.
The angles X and r that describe the shape of the wedge are de-
termined by the angles of soil-metal friction and internal soil fric-
tion. Soil-metal friction acts on the back side of the wedge (soil-tool
interface), since the wedge moves up the tool. The adhesive com-
ponent of equation 47 is small compared to the friction component
so that the principal stress acting on the back of the wedge can be
assumed to act in the direction of DR, Since the principal stress is
compressive, it will be algebraically the smallest principal stress; as
shown in figure 42, the shear surface will be inclined \~T"^~£~)
to the principal plane—a plane perpendicular to DR. Note that the
shear surface will be inclined i~T~2) ^^ ^^® direction of the
principal stress—the direction of DR. Thus, in figure 91, 5,
144 AGRICULTURE HA]SLDBOOK 316, U.S. DEPT. OF AGRICULTURE
tanf-^-A\
tanX = \-±-^-L /75y
cos 8 ' ^'^^
where X = angle between the direction of travel and sides
of the wedge in a horizontal plane.
WING
DRCOSS cos dp ^
SR COS 6
p cos op
¡c COS)ö
F cos ß
(A) (B)
FIGURE 92^.—Forces and their location while acting on a wedge of soil formed
by a narrow vertical tool: A, Horizontal plane; B, vertical plane. (Payne,
Jour. Agr. Engin. Res. {329 ).)
146 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
SIDE OF WEDGE
Sc SIN ß
(A) (B)
tool and soil parameters was used in an attempt to verify the me-
chanics.
Equations 72 and 73 may be used to predict the draft of a narrow
vertical tool at the instant of failure of the shear surfaces; hence,
the predicted draft value will be a maximum. As the tool continues
to advance, the draft should be at some lower value until another
failure state has been built up. Numerous researchers have noted
variations in measured drafts, which may be explained by a series of
shear failures as the tool advances. Some measure of the variation
should be taken into account. An analysis of the results showed
that the draft could be assumed to be nearly proportioned to cohe-
sion. Since Kîohesion ranged from a low of Cr to a high of È/, an
estimate of the minimum draft could be calculated from the relation
Y
95% œNFIDENCE LIMITS
o
500
tr, FITTED REGRESSION-
o (bxY =0.7516) /
o
Lu
400
fe
o REGRESSION
UJ
a: (b = l)
o. 300
ü.
o
<
UJ 200
Ü
I-
UJ
2 100
(A) (B)
FIGURE 95.—Soil cutting as influenced by the height of lift of the cutter and
depth of operation.
AP ^
V \
JC
<
Î
\
'h
(A) (B)
FIGURE 96.—A, Soil movement caused by a thin vertical cutter; B, relation of
cutting force to depth of operation for a vertical cutter. (Kostritsyn {2S0 ).)
(A) (B)
FIGURE 97.—Forces on and shape of two soU cutters. (Kostritsyn (230).)
26.5
S (mm)
i ï
y4 4-^
y ^2
s / /
b" b' b <
] '
maximum occurs when point a in figure 100 reaches the widest part
of the wedge section of the cutter. The maximum deformation is
shown in figure 100 in the triangle Wh'' ' where geometry indicates
that the angle VW ' is (a/2 + 8) so that the length Vh is given
by the equation
J-^max — (86)
^ cos {aß + 8)
where S = width of cutter,
8 = angle of soil-metal friction,
a = wedge angle of cutter,
Lmaw = maximum deformation.
The soil deformation along the wedge will vary from zero at the tip
to the maximum shown in equation 86, so that average soil deforma-
tion Lo can be calculated by the relation
0 + Ln S
Lo = (87)
^ COS (a/2 + 8) •
Equation 87 applies to the deformation caused by the wedge portion
of a cutter. Kostritsyn reasoned that no additional deformation
occurs along the sides of the cutter, and the average deformation
will be constant and numerically equal to that given for half of the
maximum deformation occurring on one side of the cutter and ex-
pressed by equation 86.
Kostritsyn argued that a relation must exist between soil deforma-
tion, specific pressure, and specific resistance. In reality, specific
pressure and specific resistance are the normal stresses acting be-
tween the soil and the cutter. The value of specific resistance is the
stress required to cause a given deformation. Kostritsyn recognized
that such behavior is stress-strain be^havior (uniaxial compressive
stress-longitudinal strain), but he also recognized that no suitable
relation existed. Such a relation can be represented by a simple be-
SOIL DYNAMICS IN TILLAGE AND TRACTION 155
hayior equation that would define dynamic parameters of the soil.
With the direction of strain movement specified as discussed (the
behavior output of the equation) the mechanics of cutting could be
developed in a manner similar to that used by Soehne, for example.
Recognizing the behavior equations involved and specifying the be-
havior outputs establishes a method by which to locate and orient the
forces involved. Kostritsyn, however, by analyzing the cutting re-
action, has worked backwards and shown the need for a specific
simple behavior equation. The difference that was discussed earlier
between the methods for establishing a mechanics thus should be
apparent.
Kostritsyn did not study a stress-strain relation directly, but rather
indirectly through his mechanics. For a given cutter he could cal-
culate Lo for the respective K values from equations 86 and 87. He
could evaluate experimentally the magnitude of the two K values
from equation 85 by using two cutters with different shapes. An
example of the relation between Ki and K^, and their respective
average deformations is shown in figure 101 for one soil in one
condition.
23.2 Lo(mm)-K2
Kei = K2 ^. (89)
cos a/2 ^ '
With equations 88 and 89, the elastic and plastic components of
stress can be calculated from the values of Ki and ^g. Figure 102
shows the respective values plotted against deformation, as deter-
mined by equation 87. Kostritsj^n determined the relation shown in
figure 102 for several soils in various conditions. He concluded that
the general shape of the curves was the same for all soils and then
proceeded to obtain a mathematical expression for the curves. He
noted that the relation between K^i and LQ was very close to that of
an equilateral hyperbola so he used the equation
10 II 12
( mm)
(A) (B)
the cutter to move from A to B along the projected path (fig. 103,
A), several of the large particles would have to be sliced and sep-
arated. Contrast this action to the actions implied in Kostritsyn's
mechanics for cutting, where particles presumably smaller than the
cutter are merely displaced but not sliced. A possible fourth term
160 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
could thus be included in equation 81. This term might be extremely
useful when evaluating the cutting resistance of materials in the soil,
such as roots.
Figure 103 also shows the effect a boundary condition might have
on cutting. If a pile of coal were being shoveled from the top
(a situation similar to fig. 103, A) rather than from a smooth floor
(a situation similar to fig. 103, 5), the force to push the shovel into
the pile would obviously be different. In the latter case, the bound-
ary condition becomes orderly so that cutting is not required.
Shoveling from the top requires deforming the aggregate of coal in
an action similar to that represented by Kostritsyn's mechanics. The
individual aggregates are displaced but not cut. Presumably an
action could occur wherein the coal aggregates themselves would be
severed so that an additional force would be required. Cutting per
se is thus simply envisioned and easily defined, but its involvement
with other actions compound and confuse the practical application
of cutting. Getzlaff ( U2 ) has measured the effect of stones on the
draft of plows and it appears that cutting or displacing rigid bodies
can require considerable force.
4.3.4 Conclusions
The examples of partial mechanics discussed in section 4.3 are the
only examples presently available. They are restricted, occasionally
based on questionable assumptions, and, as experimental evidence
shows, not satisfactorily accurate. These shortcomings are acknowl-
edged but do not detract from the fine efforts that have been made by
Kostritsyn, Soehne, Kawamura, and Payne. Their analytical con-
tributions have demonstrated the methods for developing a mechanics
that integrates the soil and tool into a system that can be analyzed.
Demonstrating that the actions involved m tillage can be represented
by rigorous mathematical treatment is more significant than the
practical usefulness of their mechanics. Even though the examples
discussed here are not a complete solution, the scholarly approaches
must be acknowledged as a break from traditional methods. Follow-
ing the principles illustrated will lead to the development of a com-
plete mechanics that will provide an understanding of soil-tillage tool
reactions and that will ultimately have practical usefulness.
4.4 Soil Behavior in Simplified Systems
Actual examples of soil-tillage tool mechanics have just been
discussed. In each example, the mechanics was based on one or more
behavior equations. The mechanics combined the active contribu-
tion of each behavior to the total action so that the whole reaction of
the soil was represented. In chapters 2 and 3, a number of behavior
equations that have been identified were discussed. In equation 18,
two soil parameters 0 and (f) were identified, and ample evidence was
cited to show their nature and importance. Not all behaviors have
been so well studied.
One principle for developing a mechanics is to identify the various
forms of behavior that are present. Each identified form of behavior
requires an accurate descriptive equation. One method for develop-
ing such equations is to separate the various forms of behavior that
occur in the soil and study each behavior individually. Examination
SOIL DYNAMICS IK TILLAGE AND TRACTION 161
moisture, but it may also include other forces such as magnetic at-
tractions that might occur between the two materials. Some ano-
malies that have been encountered in this assumption were discussed
in chapters 2 and 3, and apparently additional work will be required
to obtain a clear picture of the importance of the moisture suction
term. A consideration of the attraction between the soil and the dia-
phragm of a pressure transducer shows that no deflection of the
transducer occurs regardless of the magnitude of the attraction.
Thus, the normal load detected by the pressure transducer does not
and cannot measure this attraction. The suction in the moisture film
may be measured by tensiometers, but they would have to be inserted
into the sliding surface and have all of the characteristics of the
sliding surface to provide reliable measurements. Formerly these
instruments could be utilized only at low moisture suctions, but recent
improvements have extended their range considerably ( 365 ). A
major difficulty in dynamic sliding actions is the lag time inherent
in tensiometers.
The term F must represent the actual area over which the water
films are effective. This area might be observed when glass models
are used, but the area of the films within the soil is normally an un-
known quantity. The extent to which approximations of the area,
as might be deduced from measurements of bulk density, void ratio,
and moisture content, can meet this need is not encouraging. When
the entire normal load is applied by means of adhesion (sec. 2.9.3),
the magnitude of the suction is known, the area of contact of the
film is known, and the sliding force can be measured. Under such
conditions the coefficient of friction ft' can be computed in the same
manner as when the load is applied by a mechanical load.
In practice, both mechanical and suction loads are operating on
the soil so that the coefficient determined by equation 29 is an ap-
parent coefficient and wdll be identified as />(.". It may be visualized
as having the form
R = ¡x'N.^-ixfN, = />t'W„ (96)
where Ni — the load due to mechanical force,
N2 — the load due to moisture suction,
/x'' = the apparent coefficient of soil-metal friction.
When N2 cannot be measured, the coefficient is determined on the
basis of the applied load rather than the effective load, and the net
effect of the second term of the equation, /xW^, is lumped into /x".
The physical interpretation of this action has been to assume that an
increase in the sliding force is due to an increase in the apparent
coefficient of sliding friction /x'', and not due to an increase in the
normal load (by factor N2) which results from adhesion. That the
components of ft'' are not precisely known does not render equation
96 useless. The apparent coefficient />t" has been found to be a very
useful soil parameter, but methods for determining /x' should be
improved.
Equation 96 implies that /x' as defined by equation 29 is not an in-
dependent parameter but is a composite parameter. As was dis-
cussed in section 3.2, independent parameters may become composite
parameters as additional knowledge is obtained. Furthermore,
164 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
8 12 16 20 24 28 16 20 24 28
SOIL MOISTURE {%)
(A) (B) (0
length of the sliding path in sand A, loam B, and clay soil G, The
increase in /x' resulting from an increase in the sliding distance was
greatest at low moisture contents. Under the wettest conditions,
the lesser influence of the length of the sliding path on the coefficient
of soil-metal friction for steel was probably due to the low strength
of the soil. When wet, complete puddling presumably occurred with
the slightest movement and no further change was possible. Ad-
hesion was probably not altered to any appreciable extent with ad-
ditional movement.
In contrast to a sliding surface of steel, rubber sliding over the
soil presented a different relation. As shown in figure 106, little
influence of the length of the sliding path appears in either sand A
or loam soil B. In clay soil 0^ however, the coefficient increased
0.5- 0.9-1
0.5-
0.4- O8
^04
0.3 03 0.7 H
0.2-1 0.6
02
O.l 0.5
0.1- o = 30 cm
A = 60 cm 04
0 -T r- 0
2 3 4 6 10 14 18 22 26 8 12 16 20 24 28
SOIL MOISTURE (%)
(A) (B) (C)
E = X fi'iPiFi. (98)
<=j
Sliding resistance has not been calculated in this fashion and it can-
not be determined from studies utilizing simple sliders. The distri-
bution of normal pressure might be determined by means of pressure
transducers. If pressure transducers were imbedded in a tool as
shown in figure 107, the pressure at each of the small segments of
interest could be determined.
Mayauskas ( £95 ) used simple pressure transducers to determine
the normal distribution along plowshares during actual field opera-
tions. The results are discussed in section 6.4.1.
0.6 -
""■""•- LOAM
0.4
^"""-^^ CLAY
1 1 1 1
LOAD (Kg/cm^)
FIGURE 109.—The influence of the normal load on the coeflScient of sliding
friction. {Vetrov (465).)
depend on the soil materials that are deposited along the tillage tool
surface. These coating materials might be waxes, moisture, fine clay
particles, or other materials in soil. Surface coatings would account
for the change in jx^ ; however, /¿' might be expected to remain fairly
constant after equilibrium is established on the surface. After a
plow travels far enough to coat the entire surface, a uniform value
of /x' might be found for the entire surface. In a final analysis, both
the soil surface and the metal surface may change.
Physical reality must accompany raathematical representation of
pressure distributions along sliding surfaces. The few pressure
measurements that have been made along the sliding surfaces of tools
indicate that wide variations may occur. If detailed studies are to
be realistic, cognizance must be taken of interactions. Values of /x'
must be determined for the entire range of pressures distributed over
sliding surfaces. The extent to which this type of information may
alter the concepts of theory or design is not known. As an example,
it has been suggested ( 106 ) that a uniform pressure is required to
produce a uniform acceleration of the soil. Actually, a uniform
movement of soil across the face of the tool is desired ; and this may
conceivably be obtained with different normal pressures along the
surface of the tool.
Let us imagine a distribution of pressure that might occur along
a plow surface when shear failure occurs as visualized in the Nichols
model (fig. 110). If all blocks of soil are to be accelerated with a
uniform force, the resistance to movement of all blocks must be
the same. This is not the case, since the shear force along the bound-
ary between block A and the soil mass M is very large compared to
that between any of the other blocks of soil. Theory then must
SOIL DYNAMICS IN TILLAGE AND TRACTION 169
recoenize that the forces on the share of a plow are different from
those on other parts. Equation 107 (sec. 4.4.1.3) incorporates this
principle by keeping forces on the share separate from those ot the
sliding surface. Little or no twisting of the furrow slice occurs at
the forward part of the plow, but as the individual blocks progress
up along the plow surface the forces on the plow vary considerably
Indeed, the twisting action of a moldboard may pull the blocks apart
so that gravitational forces actually cause overhanging segments to
be placed in tension, as shown in figure 111. Obviously, the pressure
on the surface of the tool goes to zero where the cracks appear.
There seems to be little value in making additional theoretical
ASCENDING ANGLE - ^
SLIDING ANGLE - S
HORIZONTAL LINE
of the sliding surface and the path of the sliding particle. A sec-
ond angle, the ascending angle i//, was identified. It was measured
from a horizontal line in the direction of travel and the path of the
sliding particle. This angle might be utilized to compute the work
done against gravity.
A number of factors, such as the frictional and adhesive character-
istics of interface systems, may be expected to affect the location of
the sliding path. With the exception of the speed and the shape of
SOIL DYNAMICS IN TILLAGE AND TRACTION 171
the tool, few of these have ever been studied. As shown in figure
113, both the particle size and the angle of inclination of a plane tool
surface influence the angles of ascent and sliding. A knowledge of
90
ASCENDING X X
80 SLIDING
70
60
^ 50
UJ
á 40
z
<
30
20
10
0
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
DIAMETER OF SOIL PARTICLE (mm)
FIGURE 113.—Effect of size of the soil particle and inclination of the tool on
the sliding and ascending angles. (Kaburaki and Kisu, Kanto-Tosan Agr.
Expt. Sta. Jour. {20Jf).)
the factors that govern the movement of soil particles will be helpful
in the design of tools, since it becomes important to be able to direct
the sliding of soil along predetermined paths. These paths may-
direct the movement of the soil so that a minimum energy may be
required for the movement or so that shearing strains will break up
the soil. One of the more generally recognized directive actions is
the inversion or movement of soil to some specific location at the
boundary of a tool.
4.4.1.3 Mechanics for Draff force of Sliding Actions
The foregoing material has been presented to describe soil be-
havior in sliding. The failure zone is predetermined by the interface,
since any coupling between the slider and the soil results in shear
or soil-soil sliding rather than soil-metal sliding. In the sliding ac-
tions that have been discussed, only fiat surfaces have been considered.
The research that was reported was directed mainly toward deter-
mining the basic behavior equation and identifying the input and
outputs of the equation. In addition, studies were directed toward
identifying the basic parameters of the equation. The studies indi-
cate that more basic relations must be considered before the data can
172 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
be used effectively in developing a mechanics. The alteration of the
soil physical condition as a result of the sliding stresses creates a
condition in which an interaction becomes operative. The import-
ance of interactions of this type can be determined only by a more
detailed examination.
^ Some progress has been made in developing simplified mechanics
for sliding surfaces. In each instance, however, the coefficient of
sliding friction was considered to be independent of the area of
contact and adhesion. In addition, the ultimate goal of the me-
chanics was to evaluate the sliding resistance of the tool that was
being considered. The basic behavior equation for sliding (equation
29) was accepted, and suitable boundary conditions were applied so
that the inñuence of realistic force inputs could be evaluated.
The first attempt to write a simple raechanics for soil sliding on a
curved surface was by Doner and Nichols {106). They were con-
cerned with the action of the curved surface of a moldboard plow on
the forces on the sliding interface. (3rientation of the surface in
some direction other than horizontal would alter the normal load
because of gravity. The forces to be expected on a uniformly curved
surface and on a variable curved surface w^ere determined. They
considered that the force normal to the sliding interface of a co-
hesionless soil and the curved surface was determined by three dis-
tinct components: the weight of the soil, the acceleration force, and
a buckling force which was a component of the tangential force that
causes the soil to slide.
The first of these three forces—the weight of the soil on an area, A
—IS a function of the inclination as shown in figure 114. The force
due to the weight of the soil was calculated from the relation
Wn = AhD^ cos a (99)
in which Ah = V = Acosa ,
The third force normal to the surface results from the resistance
to motion along the surface and is called a buckling force (fig. 115).
This force might increase or decrease the normal force on the surface,
depending on the direction of the curvature. The buckling force
may be calculated at any point on a curved path from the tangential
and normal forces. At any distance S along the path, the forces
may be resolved into a tangential force Ft, and a normal force Fn^
At a more distant point S +dS, the two forces would be correspond-
ingly Ft + dFt and Fn + dFn^ Simultaneously with an increase m
the distance S, there would be a change in direction of the normal
force because of the curvature of the surface and the buckling eftect
would be
F,da = Ft^ds = Ftkds = Ftf{S)ds, (101)
FIGURE IW.—Left, The path of soil on a plow; upper right, the force on a
small increment of soil; lower right, the buckling force on soil. (Doner
and Nichols, Agr. Engin. {106).)
In incipient cases of sliding, the soil moves across the tool so slowly
that the soil on the tool acts as a rigid body which is driven through
the soil mass. Soil does not flow smoothly across the plow \yhen this
occui-s. Figure 117 shows the action of two plows shown in figure
116 in the same soil. The polytetrafluoroethylene-covered plow
FIGURE 117.- -Soil after plowing: Left, With good scouring; right, with poor
scouring.
178 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
scoured and the steel plow essentially pushed the soil aside. The
problem of handling a nonscouring so-called "push soil" is of general
interest. As a rule, the adhesion between the tool and the soil is
greater than that of the cohesion within the soil so that failure takes
place withm the soil to cause nonscouring.
Doner and Nichols ( 106 ) defined the scouring S at any point on a
sliding surface as being approximately equal to the tangential force
of the sliding added to the shear resistance of the soil F^ minus the
trictional force at the same point ¡x'Fn. They concluded from their
studies that plow curvature at the wing rather than at the share
would reduce soil sticking.
Payne and Fountaine ( 381 ) studied the mechanics of scouring
along simple surfaces and concluded that the following factors affect
the scouring of a tool in soil :
1.
^ The coefficient of soil-metal friction
2. The coefficient of soil-soil friction
3. The angle of approach of the tool
4. The soil cohesion
5. The soil adhesion
They analyzed the equilibrium conditions at the point of scouring
tor a simple system in which they considered only forces along the
sliding surface. Figure 118 shows the forces as seen from above at
OPEN FURROW
COULTER CUT
FIGURE 118.—The geometry of a simple tool at incipient scouring. (Payne
and Fountaine, Nati. Inst. Agr. Engin. {S31),)
a Situation, the sticking soil (block ABF) must be a thin layer paral-
lel to AB and the same normal stress must act on both the soil-soil
surface and the soil-metal surface. To satisfy all conditions (circle
through Ö tangent to HJ and also soil shear failure so that soil
IS separated from the soil mass) the point of tangency must also be
the point of intersection of the lines HJ and LM, as shown in figure
SOIL DYNAMICS IN TILLAGE AND TRACTION 181
120. For any other circumstances, either the metal must fail in shear
or a situation as shown in figure 119 will be in effect. Thus the re-
lations between the factors governing scouring can be determined
from the principles illustrated in figures 119 and 120.
4.4.2 Penetration
Penetration is an action that may be described by a composite be-
havior since the soil usually fails by some combination of cutting,
shearing, compacting, and ño wing (plastically) as a cutter or a
probe is forced into the soil. As was discussed in section 3.2.2.1,
failure during penetration is usually considered to occur in the im-
mediate vicinity of the tip of the probe! Penetration is thus often
termed cutting since cuttmg usually implies a localized soil failure
in the neighborhood of the cutter (sec. 4.3.3). Although they were
not discussed in section 4.3.3, Kostritsyn developed equations pre-
dicting the cutting force for cone-shaped cutters. Many so-called
penetrometers are cone-shaped so that distinguishing between cutters
and penetrometers is perhaps not realistic. In studying the behavior,
researchers have used probes, cutters, penetration, and cutting rather
loosely and interchangeably. Localized failure rather than the mode
of failure is the common basis for discussing the behavior. Intui-
tively, it appears that a penetrometer assesses soil strength, and the
inherent simplicity of the measurement contributes to its practical
usefulness. Thus, in spite of its composite nature, penetration be-
havior has been studied in some detail and has even been incorporated
into a partial mechanics.
The geometry of cutters or penetrometers is important because
of its influence on the stress distribution in the soil near the tool.
Consequently, the effect of geometry on penetration behavior has re-
ceived considerable attention. The geometry may determine whether
a tool acts as a knife-type tool that slides through the soil without
soil sticking to its surface or whether it creates a compacted body of
soil that sticks on its surface. As shown in figure 121, a compacted
mass of soil may gather on a blunt tip and move with the tool as an
intricate part of the tool. At this time, primarily soil-soil friction
is active since most sliding is between soil and soil. Even though
a point is blunted or rounded, it may have little influence on the
external appearance of the compacted soil body or core and on the
182 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
force required to move the tool in the soil. \s the shape of the tip
approaches that of the shape of the soil body, the tendency for the
soil to slide from the tool increases and ultimately the frictional
resistance along the surface of the tool is reduced to where sliding
begins. At this time, primarily soil-metal friction is active so that
the force to move the tool may be less than when soil-soil friction is
present. According to Zelenin (516), the compacted core appears
when the angle of the tip exceeds 50°.
Attempts have been made to determine whether the presence of the
compacted body of soil influences the resistance to penetration.
Data reported by the Waterways Experiment Station ( 472 ) indi-
cate that the shape of the tip of the probe may have only a small
influence on the resistance to penetration (ñg. 122).
Ü 200
abc
fe 100
V)
UJ I5»C0NE
oc CONE
a: 2in FLAT
UJ ■ 45» CONE
Ui
O 10 ,<^>''n CONE
45» CONE
^lin HEMISPHERE de f
37* 30' GONE
a.
UJ
§
10 100 10 10
RESISTANCE TO PENETRATION (psi) PENETRATION (¡n) PENETRATION (In)
Figure 122, Ä compares a 30° right circular cone with flat and
hemisphere-shaped tips. The measurements were made on slowly
moving penetrometers and, with the exception of the cone, a com-
pacted body should have been present. The data indicate little dif-
ference due to the presence of such a body.
Figure 122, B shows the influence of cone angles of impact
penetrometers. The impact measurement was made on a 1-inch cone,
driven with one blow from a 1^-pound hammer dropped 4 inches.
Figure 122, 0 shows the relation between a 30° static cone penetro-
meter and a 45° impact cone penetrometer where 5 inch-pound of
energy (see curve a) and 8 inch-pound of energy (curve i) were
used. For the 15° cone, 5 inch-pound (curve c)^ 7.5 inch-pound
(curve d), 8 inch-pound (curve e), and 30 inch-pound (curve /)
were used. At the higher energies and lower cone angles, deeper
penetration resulted; but in all relations the slope was constant.
SOIL DYNAMICS IN TILLAGE AND TRACTION 183
^1 = cos {a/2
T^^X^'
-^ oy (111)
where Z) = a constant,
8 = angle of soil-metal friction.
Equations 110 and 111 can be substituted in equation 85 and, recalling
that fji = tan 8, the equation becomes
UJ
o
Q:
o
e>
z
3
Ü
10 20 30 40 50 60 70 80 90 100
an
FIGURE 123.—Computed cutting force as a function of wedge angle for a con-
stant thickness of cutter at 3 angles of soil friction. (Kostritsyn {2^0),)
45°50' and 29° are angles of soil-soil friction i//, and 40°30' is the
angle of soil-metal friction 8. The data indicate that a minimum
force occurs at a wedge angle at approximately 45° regardless of the
type of the angle (i// or 8). Thus, even if soil sticks to the cutter so
that soil-soil friction becomes active, nearly the same optimum wedge
angle results. The 45° friction angle shown in figure 123 is a typical
soil-soil friction angle, according to Kostritsyn. He reported ex-
perimental data that tended to confirm his calculated results. A
minimum cutting force would thus appear to result regardless of
the magnitude of the cutting force, and this minimum occurs at a
wedge angle of approximately 45°.
Perhaps the most thorough study of the influence of the geometry
of penetrometers on resistance to penetration was made with di-
mensional analysis techniques by Kondner and associates {85-87,
223-227), While Kondner envisioned his tools as model footings,
their cross-sectional area was less than 3 square inches. Dimensional
analysis gives the functional relationship between tool and soil
variables as
SOIL DYNAMICS IN TILLAGE AND TRACTION 185
where QG = penetration,
t — time,
F — total force,
G = perimeter of tool,
A = cross-sectional solid area,
0 — tip angle of tool,
T = maximum unconfined compressive strength of soil,
7) = viscocity of soil.
The dimensionless terms were interpreted by Kondner to reflect
certain physical characteristics of the system. The dimensionless
terms in equation 113 reflect, respectively, penetration (the dependent
W G^ .
variable) ^, strength ratio of soil -j-^ shape effect of tool ^, tip
characteristics Ö, and rate of penetration as influenced by viscous
creep of the soil —. By allowing time for equilibrium during
a static test and by maintaining a fixed tip angle, the last two terms
in equation 113 are constant so that the relation simplifies to
0.100
0.075
^ = CONSTANT = G
x|^ 0.050-
0.025
g^g^^l- ^^ Figure 125 shows the data plotted against the shape
factor -Ç. The data indicate that greater penetration occured for a
load as the ratio of perimeter/area approached a minimum. The
consistent collapse of data about lines of constant terms (terms m
equation 113) confirmed the implied functional relation. Kondner
attempted to investigate the viscous creep term ^, which is import-
ant in a dynamic situation such as the vibratory cutting of soil. He
made some progress but concluded that the relation was complex
because the viscosity appears to be a function of the stress level.
For static situations, however, the term could be considered essentially
constant if the test were conducted slowly with time allowed tor
equilibrium. Kondner thus concluded that equation 113 could be
reasonably simplified to
(115)
-y^
where h = constant,
d = constant,
S = constant.
The magnitudes oih, d, and S are determined by the specific rela-
tions of the intercepts and slopes to the strength term. Figure 128
shows the results of the varying tip angle upon — when expressed
G
5r;c, C2 C3 C4 c, C«C
'6 ^7
^= CONSTANT =C
Í2 I
-^ = /(lOÖ)-^/^ (IIT)
G
where / = intercept,
Q — angle in radians,
m = slope.
The relation between / and m and the strength term was not as
simple as for the data shown in figure 127. Equation 117 could not
be made as descriptive as equation 116, since the intercept and slope
are not constants. Equations 116 and 117 are not presently useful
since they are restricted to the soil conditions Kondner mvestigated.
Furthermore, equation 117 is restricted in application to circular
shapes where c^/A is a minimum and equation 116 applies when Ö
has a constant value. If equations 116 and 117 could be combined
into one expression, i\\<^ latter restriction would be eliminated.
The most important contribution by Kondner is the accuracy ot
the combination of variables indicated in equation 113. They per-
mitted him to consistently collapse data to an acceptable degree.
Furthermore, he demonstrated techniques for actually determining
the behavior equation describing the composite behavior, penetration.
The geometry of cutters must be studied further to reconcile the
differences that appear to exist. As an example, occurrence of a
minimum cutting force as determined by Kostritsyn is not reflected
in the data of either the Waterways Experiment Station or Kondner.
The occurrence of minima such as those due to the circular shape
and the cutting angle must be sought and verified because of their
importance in the design of practical tillage tools.
Zelenin {51S) has developed empirical relations between the
draft force of a cutting tool and physical conditions of the soil
as measured by a penetrometer. The relation was developed to the
point Avhere it constituted a partial mechanics. Zelenin conducted
a large number of experiments in which he used horizontal cutters
of the type shown in figure 77 and measured the draft and depth ot
cutting. The size of the cutting tool and depth of operation were
such that pure cutting as defined by Kostritsyn (sec. 4.3.2) was not
the only quantity being measured. Thus, types of soil taiLure m
addition to pure cutting were involved. Zelenin observed that the
draft and depth were parabolically related according to the relation
p = M^ (118)
where P - cutting force (draft) of a horizontal blade,
h — coefficient of soil resistance,
h — depth of operation,
fi = coefficient.
Based on a wide range of soil and moisture conditions, the value for
n was found to be approximately a constant whose value was 1.35.
Zelenin further observed that the coefficient Jc was directly propor-
tional to 0, the number of blows of an impact penetrometer (sec.
190 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
Measured Computed
draft Difference
draft
(kilograms)
Kilograms Percent
305 290 -4.9
260 258 - .8
620 580 -6.5
1400 1,470 + 5.0
1 200 1,130 -6.0
700 747 + 6.7
130 shows compacted soil bodies that were formed on the surface of
a tool at two speeds. The photographs are frames from a movie
film of the tool operating in a clay soil. One side of the soil bin was
Pioi-RE 130.—Soil body formed at two tool .speeds: Left. O.Ol.'i in.p.Ii. ; right,
O.r» m.p.h.
FIGURE 131.—A soil body formed on a subsoiler during tillage of a dry clay
soil. (Nichols and Reaves, Agr. Engin. ( 322 ).)
the geometry of the tool, the forces on the tool may not remain fixed.
In most cases the wear of a tool beyond the point where a refined
degree of polish promotes scouring results in an undesirable situ-
ation.
An analytical study of the alteration in forces due to the change
of geometry by wear was made by Gavrilov and Koruschkin ( HO).
Based on the observation that wear occurs along the underside of the
tool, the angle of change in sliarpness of the tool a was visualized as
shown in figure 132. As wear increases on the underside of the edge,
the clearance angle y decreases to the point where it could become
negative and in fact become an angle of approach. When it becomes
an angle of approach, there is no longer clearance under the tip and
196 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
X-^
FURROW BOTTOM
(A)
a larger area of the tool is in contact with the soil. Under severe
wear conditions, the angle can increase to the relative magnitude
shown m figure 132, B so that a normal force B develops on the front
oí the worn surface Ä0. Soil-metal friction changes the direction
ot the resultant force to Ä, = A/cos S. With a forward movement
ot the tool, a horizontal resistance P^ develops in front of the tool.
Ihis resistance is related to both Ri and R in the form
Pi = Risin (8 + 7), (124)
orPi=R sin (^\ (125)
\COSÔ/
Equation 125 can also be expressed in the form
Pi - R (tan 8 cos 7 +sin y). (126)
Since tan 8 represents the coefficient of soil-metal friction, it should
be possible to determine relations between P^ and y for a fixed value
of 8. By equating the first derivative of equation 126 to zero, a
relation can be obtained where
P' — R (cos y - tan 8 sin 7) = 0. (127)
When P' becomes a maximum (fig. 133), it is possible to establish a
relation between 8 and y having the form
cosy
tan 8 (128)
sin y tan y*
Thus, when Pi becomes a maximum, there is a reciprocal relation be-
tween the angle of sliding friction 8 and the approach angle y. In
practical situations, this relation might determine the final shape of
the wear pattern at equilibrium.
The frontal pressure that is applied to the underside of the tool
for a unit of width would cause a moment M^ about the point O
which would tend to lift the tool out of the ground (fig. 132).
SOIL DYNAMICS IN TILLAGE AND TRACTION 197
150
rn
FIGURE 133.—Relation of the clearance angle 7 and the force P^ when 5 is
25°. (Gavrilov and Koruschkin, Selkhozmashina {I4O).)
(129)
where Ri = ^,
cos Ô
L =
cos (8 + 7) *
On a small increment of the surface along AC,
R xdx (130)
dM = coso COS (8 + 7) '
and when x = B cos y,
R B^ cos y (131)
M = 2 cos^ 8 - sin ^ 8 tan y
2 450 550
15 0 "W/ZÁ
1 240 415
J//A
SOURCE: Zelenin {515).
where the surface influence did not exist. While these studies served
a purpose, the need remains to examine the soil-tool relations near
the surface where the variable boundary conditions exist.
Zelenin attempted to explore the soil confinement along a vertical
boundary that was essentially an open furrow wall ( 515 ). ^ata m
table 13 show that the cutting resistance of the soil adjacent to the
wall was considerably less than it was farther from the wall where
the confinement was greater. Eegardless of the thickness, shape, or
angle of operation of a tool, enough soil must be displaced to permit
patsage. ^Consequently, for the conditions m table 13, c^tsjnade
m excess of 35 centimeters from the wall required an ultimate cutting
force of about 75 kilograms. The actual distance reflects the intlu-
ence of geometric characteristics of the soil and the tool boundary
and dyntmic behavior patterns of the soil. The distance from the
open wall where the ultimate draft resistance is reached reflects the
point where the total displacement or strain required to cause ade-
quate failure is absorbed within the mass of the soil. At lesser dis-
tances from the wall, the soil is probably detached and moved into
an open furrow as a rigid body in order to provide ™om for pas-
sage of the tool; hence, the forces required for displacement are
reduced. Data are not available of side.forces on tools operating
in these conditions but they should provide information that would
be of assistance in evaluating the influence of geometry in such a
^^Nothini in Sie previous discussion implies that the specific strength
of the soil was cliianged by the different soil-tool boundary condi-
tions The change in draft of the tool was due to the different
200 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
Kilograms
5 2.8
10 10.1
15 22.1
20 31.3
25 47.6
30 73.3
40 75.0
1 Cutting depth, 15 cm.
SOURCE : Zelenin {515).
URROW WALL
FIGURE 135.~Plan view of different soil-tool systems in which the soil reaction
IS essentially the same. (Zelenin {515),)
SOIL DYNAMICS IN TILLAGE AND TRACTION 201
95.9 Kg
70.0 Kg
- ^^
" -45» -30' 0^ +30* +45*
INCLINATION
1.5
UJ
o
z
1.0
</)
UJ
û:
o
0.5
z>
o
-L
10 20 30 40 50 60 70 80 90
CUTTING ANGLE ß (*>)
FIGURE 137.-Effect of the side angle on the draft of a simple inclined tool
(Kaburaki and Kisu, Kanto-Tosan Agr. Expt. Sta. Jour. Uö^^ )
SOIL DYNAMICS IN TILLAGE AND TRACTION 203
essentially doubled when the lift angle was increased from 20° to 90°.
Increases in the side angle ß decreased draft until an angle ot ap-
proximately 40° to 50° was attained. After that pomt, dratt be-
came essentially constant for each value of a. In no case did the
decrease in draft exceed 25 percent. For one special case, where
a = ß = 45°, a parabolic increase in draft was found to be due to
an increase in the width of cut regardless of the depth of cut. This
relation has an important effect on the design and use of tools when
optimum draft relations are of interest. The results of the influence
of other orientations on the draft of a tool are shown m table 3b,
where increases in the angle of approach result m increases m the
draft force. This orientation is easily described m a simple tool sys-
tem, but we shall see that much research is needed with regard to the
orientation of tools having more complex shapes.
When the main tillage action is cutting, the size of the isolated soil
mass is determined by the size of cut of the tool. In a number ot
cases, however, the final projected area of disturbed soil is not the
same as the projected area of the tool. Because of this, tools may be
located and oriented so that their sphere of influence includes all
of the area to be tilled even though the tools do not intercept all ot
the periphery. Kostritsyn ( 230 ) reported data of Dalm and Pav-
lov, who measured the area of soil disturbance of small cylindrical
tines (table 15). The data show that either soil bodies must form on
tools or the arching effect in the soil results in a disturbed zone ot
soil considerably larger than the projected area of the tool.
TABLE 15.—Sphere of influence of cylindrical tiries in soil
Diameter of tine Width of sphere of
(millimeters) influence
Millimeters
7.5 150
12.4 176
17.4 193
28.9 210
SOURCE : Kostritsyn ( ¡
(A) (B)
mwrnim/u
RUPTURE ZONE
TOOL
SOIL BODY
SPEED
0.4 m/sec
—— 1.6 m/sec
gravity /S' of a unit of the plow furrow slice when plowing was done
on a 15° slope with a right-hand plow at two speeds. The initial
location S is shown at the center of the direction rays. The final
location of the center of gravity of the plow furrow is shown by S^j
where n is the numbered direction in which the plow traveled. The
experiment demonstrates a method by which the soil reaction was
described in a simple quantitative manner. Since the minimum
movement of the centroid of the soil mass represents the minimum
expenditure of energy commensurate with the work accomplished,
perhaps this technique can be used to partly evaluate the performance
of plows.
SOIL DYNAMICS IN TILLAGE AND TRACTION 207
TOP VIEW
FRONT VIEW
/\
increased still further, the draft increased until the point d = '2.6t
was reached and the cutters were acting independently.
Zelenin ( 515 ) repeated experiments of this type at several depths
and found a similar relation (fig. 143). At the shallow depths, the
DEPTH (cm)
18
5 10 15 20 25
DISTANCE BETWEEN TOOLS (cm)
5.1 Introduction
Tillage is the manipulation of soil by mechanical forces. The
purpose of tillage tool design is to create a mechanical system, that is,
a tillage machine or a series of machines capable of controllmg the
applied forces in order to achieve a desired soil condition As a
matter of definition, a tillage tool will be considered a single soil-
working element whereas a tillage implement or machine will be
considered a group of soil-working elements. A tillage implement
or machine will include the frame, wheels, or other structural units
that are needed for guidance and support. Although tillage is
nearly always effected with an implement, the emphasis here will be
on the design of tillage tools rather than implements.
The pressing need for design information has demanded that
methods for design be developed. In fact, the need is so great that
qualitative procedures have been and still are widely used. The
qualitative procedures have often been based on art rather than
science (121, 269). That these procedures must be changed it
progress is to be made in tillage tool design is clearly demonstrated
by the history of tillage tools. , • -^ .
Basic tools such as the forked stick date back into antiquity; yet,
they are still found in their original form in many parts of the
world. Even in more advanced societies, today, the moldboard plow
is designed by empirical methods. Generally, these empirical meth-
ods are trial-and-error attempts; the tool is varied m some manner
and acceptable designs are identified when the resulting soil condi-
tion is adjudged to be satisfactory. Quantitative descriptions or
representations of the final soil condition are seldom used and, m
addition, the forces required to move the tool are frequently not
quantitatively assessed. Generally, no effort is made to describe the
reaction of the soil. Consequently, design today merely accepts
what occurs; it does not control what occurs. Thus, even though the
need for design is great, design in the true sense of the word is not
accomplished and probably will not be accomplished until quantita-
tive information is available. .
To illustrate the pressing need for design information, consider
the economic possibilities of the results of better design. In the
United States, more than 250 billion tons of soil are estimated to be
stirred or turned each year (268). To plow this soil once requires
500 million gallons of gasoline costing $105 million. If proper de-
sign could decrease the draft of the plow only 1 percent, a savings
in direct operating cost of $1 million per plowing would result. It
soil manipulation can be controlled by proper design so that subse-
quent operations may be minimized or even eliminated, additional
211
212 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
INITIAL
SOIL CONDITION FINAL
SOIL CONDITION
TOOL SHAPE
^í;;íííí;.soiLí;íí;ir;rí
»^;',MANIPULATION;í;;
als 01 m 0^>i
■ ■ ■—I i »—I »
(A)
71
SLIT LIGHT
PROJECTOR
CAMERA
(A)
i*P^o
I 1 \f\oc///y/' 4^.0 y
\j^/v/ y\y
Jl^/y^y / X
■ú^^i^ /i A. X
^yyx^ / /
' ^ ^^^^ //
1 ^ "^y^ \/^ X
^^^ \ \/
\ \*'5LÍ!^^/^
\ V<fc■^/l^r\o
b
^ 1
\ ^ TX:
VPIOJ
(B)
FIGURE 149.—Graphical representation of plow macroshape : A, Projections
into a vertical plane of vertical contour lines formed by the intersection of
the moldboard surface and a vertical slit of light; B, projections into a
horizontal plane of horizontal contour lines formed by the intersection of
the moldboard surface and a horizontal slit of light. ( Soehne, Grundlagen
der Landtechnik ( ^ö^ ).)
SOIL DYNAMICS IN TILLAGE AND TRACTION 225
One of the first attempts to relate shape to newly created soil con-
ditions (equation 133) was made by Ashby (19), He utilized
graphical representations from which he defined parameters ot the
shape of plow bottoms. He attempted to correlate the shape para-
meters with observations of plow performance. His measure ot
performance was primarily the covering of plant residues during
actual plowing. Ashby defined a slope coefficient that was identified
with the "full cut section" of a plow (fig. 150). This section is
FURROW WALL
"I
D'
>
7
i t
(A) (B)
FIGURE 151—A, Profile of fuU cut section of plow; B, profile of twist section
of plow. (Ashby (19).)
' 226 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
as a first attempt. One hopes that more than covering can be in-
cluded as a measure of the newly created soil condition. In addi-
tion, a quantitative description of the covering should be developed
to replace the standardized rating system. Although equation 141
is inadequate and probably inaccurate in terms of design needs for
today, its weakness should serve as an impetus to seek improvement.
In principle this rudimentary equation demonstrates that design
equations can be developed.
Soehne ( 402^ ^03 ) defined a number of parameters of shape de-
scriptions and attempted to relate them to plow performance. He
used the light slit technique to represent shape along grid lines 1-11
and a-m, spaced 40 millimeters apart (fig. 149). And he selected the
following angles as parameters:
= share cutting angle,
<}>10a to (^lOj - lateral directional angle of the moldboard upper
edge,
</>5a to <f)si = lateral directional angle of the horizontal shape line
5 at the intersection with vertical shape line a to ^,
8al to 8« = share intersection angle at the vertical shape line
a to h^
8a5 to 8(5 = angle of the vertical shape line a to i along the hori-
zontal shape line 5,
8A, 8B = angle of the vertical shape line at A and B.
FIGURE 152.—The sou sliding path at two speeds of plow operation. (Soehne,
Grundlagen der Landtechnik ( 1^02 ).)
hour than at 1.75 miles per hour. The path of travel is related to
the acceleration that is imparted to the soil ; this, in turn, affects the
acceleration force that contributes to draft. After examining a
number of plows, Soehne concluded that so-called high-speed plows
accelerated the soil less than did so-called standard-speed plows.
Carlson ( 60 ) reached the same conclusion from a study of two
plow shapes in which he used a digital computer to calculate the
acceleration characteristics of each shape. Even though the mathe-
matical representation of acceleration is simple and fully developed,
acceleration has not been quantitatively incorporated into design.
Nichols and Kummer ( 319 ) developed a tracing apparatus that
could provide empirical equations to describe the path of soil move-
ment over a moldboard plow (fig. 153). The principle of the ap-
paratus was similar to that of Jefferson {199)', he used a predeter-
mined physical path along which to establish plow shape. The
apparatus of Nichols and Kummer used the plow shape as the
"directrix" to determine the orientation of the generator. Kecall
that Jefferson prescribed the orientation of the generator which
thus swept out the shape.
The tracing method was based on the observation that most plow
shapes represent a section of cylindrical surface. Therefore, a se-
lected test arc may be fitted to the surface of a plow and swept across
the plow along a hypothetical path of travel. As shown in figure
153, a measurement carriage rolling on guiding rails permits the
test arc to be moved along the plow. At various increments of
travel, which can also be considered to be units of time, the rotation
SOIL DYNAMICS IN TILLAGE AND TRACTION 229
ARC
ROTATION
ANGLES.
FIGURE153.—Apparatus for tracing the path of soil movements and the shape
of moldboard plows. (After Nichols and Kummer, Agr. Engin. { 319 ).)
100 -
CALCULATED VALUE
c^--o MEASURED VALUE
20 30 40
S (cm)
FIGURE 154.—Constructed and actual paths of soil movement on moldboard
plows. (Pfost, Auburn Univ. {334).)
final soil conditions as implied by equations 132 and 133. Until such
relations are effected, quantitative design information will have to
originate from trial-and-error procedures.
Gross descriptions of shape, even though nondetailed, can be useful
for design. Qualitative descriptions such as can be developed by
the Jeffersonian technique have contributed greatly to the develop-
ment of moldboard plows. The "use of one such gross description
was reported by Kaburaki and Kisu (205), Modifications made in
the shape of moldboard-type plows have resulted in better scouring
characteristics for Japanese conditions. These developments have
led to an elliptical-shaped plow (fig. 155). The exact nature of the
superior f rictional relations was not determined ; but the tool had a
smaller width of cut, which may have produced lower normal forces.
Even though the design relations have not been established, the
elliptical description permits a convenient shape characterization.
Since the tool is symmetrical, it could also serve as a two-way plow,
as shown in figure 155.
Moldboard plows are the most widely used soil loosening and turn-
ing tools, but subsoilers require the largest draft force to move them
through the soil. Thus, it is not surprising that the force-shape re-
lation has received emphasis in subsoiler studies whereas the man-
ipulation-shape relation has received emphasis in studies of the mold-
board plow. The large draft requirement of subsoilers is the domi-
nant characteristic that determines the extent to which it can be
operated. Thus, the extent of soil breakup is of secondary import-
ance, even though it is the reason for operating the subsoiler.
Nichols and Eeaves ( S£2 ) measured the draft of a series of sub-
soilers with macroshapes that ranged from the normal straight con-
figuration to a deeply curved configuration (fig. 156). Draft was
measured in several soil conditions, and the results indicated that
SOIL DYNAMICS IN TILLAGE AND TRACTION 231
/ /Í'AW/'A*,
/^î^^^
the subsoiler with the most curve required the least draft (table 16).
In a highly compacted and cohesive soil the curved tool required
from 7 to 20 percent less draft than did the straight tool Ihis
decrease is substantial, and crude observations indicated tlmt the
resultant soil breakup was approximately the same tor all tool
shapes The curved subsoiler presented an operational ditticulty,
however, since its greater length made turning and guiding the tool
while it was in the ground difficult. Xo effort was made to describe
the shape or to relate shape to draft except in the qualitative manner
indicated in table 16. ,. , i i *,.
Improper operation can defeat the advantage of decreased draft
with a curved subsoiler. Unless the curved tool is operated at its
intended depth, all advantages of the curve may be lost. Presum-
ably, the curved subsoiler gains its advantage from the direction in
which it applies forces to the soil and the direction in which these
forces cause the soil to move (fig. 157). The advantage of the
proper use of the design ( fig. 157, A ) is lost if operation is too deep;
232 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
Sand — 0^
Sand (%) and Cecil
clay (%)_ jai
Cecil clav M 3A
Siisaiic^hanna clav - .48 .59
T^iifkin clav .48 .51
1 Summarized from 1,000 measurements.
2 Surface normally i)r()vided by manufacturers of tillage tools.
SOURCE : Nichols ( 3/6 ).
FiouBE 158.—Soil bodies formed along the sliding surface of a tool because
of minor surface Irregularities.
SOIL DYNAMICS IN TILLAGE AND TRACTION 235
them along the surface. The effect on draft performance of these
nonscouring spots depends on the area of the spots as compared to
the total surface.
Friction is greatly modified by adhesion. As discussed m section
2.9.3, adhesion results primarily from moisture films that increase
the perpendicular attractive forces between soil and the sliding
surface of a tool. Adhesion may be reduced by using a material
that does not readily wet with water. Polytetrafluoroethylene and
polyethylene, which have nonwetting characteristics, are currently
being used as coatings on moldboard plows and other tools ( 81^ 336),
The coefficient of friction of these materials can be as much as 50
percent less than that of smooth steel {93, 133), When lack of
scouring is a problem, polytetrafluoroethylene can improve scouring
so that the full draft may be reduced as much as 25 percent (table
18). Once applied to a sliding surface the material remains effective
until it is worn away by the abrasive action of soil.
1,500 0.266
/^'
0.333 0.250
3,000 .266 .333
4,500 _ .250
.266 .333 .244
SOURCE: Nichols ( n2).
the action of the air could eliminate any friction between a tool
and soil. Figure 159 shows a scheme for introducing a cushion of
air between the soil and the tool surface. The effectiveness of this
scheme is influenced by the permeability of soil to airflow, since.a
SOURCE OF
COMPRESSED AIR
'^'^^.^
^¡mmW'
the operation. Since the tool was designed to cut weed roots, shatter
of the soil and soil movement were not particularly desirable. A
simple straight tool operating at right angles to the direction o±
travel is seldom used. The blades are usually swept back at angles
of 20° to 50°. This permits self-cleaning of the blade.
Any alteration of the orientation of a sweptback tool with respect
to a horizontal plane has a secondary effect on the depth of cutting
(fig. 161). An increase in the approach angle of the tool shank ot
2:^
— ^
FIGURE 164.—Deep tillage tool with a freely rotating spinner intended to cause
greater pulverization of the soil.
been found, and the spinner increased the draft of the tool approxi-
mately 8 percent. No attempt has been made to introduce power
into the soil by driving the spinner during operation; possibly this
would have a pronounced granulating effect.
There are instances when intuition indicates that a particular
movement of soil might be accomplished more readily by the inde-
Eendent action of several shapes than by a single shape. Soil may
e inverted by a moldboard plow, but some mixing within the furrow
slice usually occurs. Where a thin surface layer is to be completely
covered, it might be expedient to use two decisive actions. The
first cuts the layer loose and deposits it in the bottom of the furrow-
while the second covers the deposit without mixing. This might be
desired when radioactive dust or some other material must be buried.
Figure 165 shows the actions accomplished by plowing a deep furrow
and using a small skimmer blade or plow to push the surface ma-
terial into the bottom of the previously created furrow. Depending
242 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(A) (B)
(C) (D)
(E) (F)
INSIDE BEVEL
400
OUTSIDE BEVEL (A)
200
OLHI-
(B)
Ö -2001-
£
_J -400 •
L-Hh
-200
-400
L. "20^ 30
disk (262). Figure 168 shows the effect of sharpening on the forces
of a disk.
Chase ( 67 ) reported that the angle of the approach edge is im-
portant. Figure 169 shows how an upper and a lower bevel on the
edge of a plane tool affected soil movement. When the bevel was on
the upper surface, a "low pressure area" caused the soil to adhere to
the surface as shown at the right in rigure 169. Soil adherence in-
creased the draft of the tool. When the bevel was on the lower side
of the surface, sticking was not observed. Chase also reported that
tools needed sharpening more frequently when the bevel was on the
top.
The forward edge of a tool, such as the share of a moldboard plow,
can cause soil compaction if it becomes blunt. Figure 170 shows how
compaction is caused through the movement of the soil. The blunt
edge shown simulates a dull share except that the edge is approxi-
mately 10 times thicker than that of a normal share. Thus, the
effect was exaggerated so that movement of soil particles along a
glass plate coated with powdered aluminum could be photographed
to indicate the effect. The blunt edge Ä caused a buildup of soil,
which in turn forced some soil to move downward and cause com-
paction at G. Soil in areas B and C moved upward into a zone of
less confinement. Vertical cracks in the bottom of a furrow of a
moldboard plow, similar to those seen at area H, have also been ob-
served with earth-moving equipment {338). Forces resulting from
a blunt edge applied to the soil in the direction of travel cause the
soil to pull apart. The number and size of the cracks depend on soil
conditions.
The compacted zone may be very thin and consequently of little
significance. In wet soils, however, the smearing action could con-
ceivably close the soil pores completely so that no air or water could
be transferred across the layer. A practical example of the detri-
mental effect of compaction was reported by workers at the Eoad
Research Laboratory {370). They found that in making small cuts
with a rotary tiller the cutting blade had to pass through an area
SOIL DYNAMICS IN TILLAGE AND TRACTION 245
M^//¿^//
(A) (B)
because of their small area and the concentration of wear, edges tend
to change their shape rapidly. The macroshape of the entire tool
generally remains relatively unchanged as wear progresses. Edge-
shape, on the other hand, is greatly changed with wear so that wear
and edgeshape must be considered simultaneously,
^i."^^^ radical change in edgeshape that can occur in plowshares and
the change m forces required to operate the plow bottom are shown
m table 21. Change m edgeshape due to wear can produce a signifi-
cant eflect on forces on both rolling and rigid tools (fig. 162). A
New__ 266 51 66
New. 290 33 64
negative vertical force, as shown in table 21, indicates that the bot-
tom had to be held m the soil to operate at the designated depth A
positive vertical force indicates that the plow had to be held upward
to prevent it from going deeper. In normal operations the mold-
board plow IS free to float and seek its natural depth as a result of the
balance of forces. Gavrilov and Kofuschkin ( 1^0 ) determined the
coeliicient of depth variation for new and worn plowshares. They
deñned the width of chamfer as the length of the line A and C shown
Í? uPin -^1^^' ^1^^^ ^^®^ ^^^ increasing width as a criterion of wear,
iable 22 shows how the coefficient of variation increased as the width
o± chamfer increased and the average furrow depth decreased
beveral researchers have demonstrated that wear occurs rapidly
Cjayrilov and Koruschkin {UO) showed that wear increased draft
resistance as much as 30 percent (table 23), and that nearly half of
the increase had occurred after only a few acres of land had been
SOIL DYNAMICS IN TILLAGE AND TRACTION 247
TABLE 22.—Effect of the width of chamfer on the average depth of
furrow in a sandy loam soil
Width of Coefficient of
Average depth variation in
chamfer
of furrow furrow depth
(millimeters)
Centimeters Percent
0 21 8.7
9 _ - 20 9.9
12 20 11.4
15 18 16.0
SOURCE : Gavrilov and Koruschkin mO ).
plowed. Figure 172 shows how the specific resistance of new and
worn shares increased with hours of operation. The rapid increase
in total draft detected after a few hours of operating time indicates
significance of wear.
^ 0.56
g 0.54
WORN-^
u, 0.52
o
§ 0.50
CO 520
UJ
a: o 18 WORN-
0.46 t
NEW
g 044
' ' I »
£8 '^ -I I L—J
^ 0.42 18 20 22 24 26 28 30
18 20 22 24 26 28 30
TIME (hr) TIME (hr)
'^ké^Ê^ ^
disks are caused by the soil they must move because of their position
and because they tend to operate deeper than the center disks of a
gang. Data from field experiments (table 24) indicate that end disks
wear nearly three times faster than center disks {353), Position,
therefore, affects normal force and thus affects wear and changes m
tool shape.
"iSAÏl*" --.Î'??..-ASä.'^^'^S-Ä
IV^^f^' rolling is not efficient; a flattened roll tends to slide and
KÍ rÄlfT "^""fr' rr^ periodically as the roll flops over
Kuhn ( mß, m ) and Garbotz and Drees ( 138 ) studied the effect
of the shapes of blades on their ease of filling and their draft requiS
W .n ^^'' Parabolic blade (fig. 174) permitted a collapse of IS so
ttiat some pushing rather than rolling was necessary. The two in
volutes resulted m good rolling. The force required to operate th^e
Th« rLTu "V^^s^red m two soil types at one speed and one deX
Se TtabfetT '41.."^^°^'?^/? '\^ ^"^^^ «^ inclination of^the
do i^^iSfilV ,^lthough the data show no consistent trend, they
do indicate that shape can be extremely important. In the sandv
silt soil, the angle of inclination changed draft by a factor of two foî
toUe
volute AtwrÎ^^Î^:,
I and the other two blades^Y also ^^^«^^^«^
accounted i^
for«hape between
a change in-
in draft
of a factor of two In other data not reported here, a mofe ncHned
Sn.f^ a smaller draft at varying speeds and depths. No con-
sistent trend was apparent between depth and blade shape. The
SOIL DYNAMICS IN TILLAGE AND TRACTION 251
by a factor of more than 1.5 (fig. 176, B and O). In sticky soils, an
accumula,tion of soil at the top of the blade (fig. 176, Z?) may ¿re-
vent the blade from filling. If the top of the blade is tilted Kar
forward, (fig. 176, F) soil will again adhere. A good shape, similar
to that m figure 176, E, will fill completely and permit good soil
movement. A good or practical shape is really a compromise of
fnTfi "l^i^li^ed shapes that serve special purposes. When used as
an isolated specific blade for one type of soil manipulation, the
number of specialized shape characteristics may be limited, and an
actual desim may overemphasize one particular characteristic.
J^igure 177 indicates several idealized shape characteristics as-
sociated with specific soil handling functions. An integration of ap-
propriate features from the idealized shapes provides a start for
designing a blade for some particular purpose.
When soil must be transported over distances that make bulldozing
W.T™'''^ •' "r""^?,^""u. ""î"^ "'^^- L^y«^« «Í soil are generally cut
^ose by an mclmed blade and the separated soil is collected in a
wheeled device for transportation. Figure 178 shows a general ar-
X?îrV*/A '^**-"^
Wl t J. •
^1^1*^^' '^°^^ *^ ^«"««t soil' and^ transport
«1^^? problems in designing a scraper so that a full
aï(2) fh^shipe'Ä bo^wl.^'^ *'^ '^''^ ^' '^^ «"^*-^ ''^'^
Garbotz and Drees ( 138 ) studied the influence of angle of inclina-
üon and depth of cut on the draft force of large indined blades
The blades were 1,000 millimeters wide and were 450, 308, and 218
millimeters long, respectively, for blade angles of 20°, 30°, and 45°
Eesults show that an angle of approximately 30° is opt mum foi^
minimum draft (fig. 179). The minimum draft coincideT^^th a
SOIL DYNAMICS IN TILLAGE AND TRACTION 253
o
6000
3
^4000 100 mm
p
Ö
z 2000
I
20* 30*» 45*»
BLADE ANGLE (•)
5.3.3 Conclusions
Macrosurface and edgeshape are important in the design of soil
loosening and turning tools and soil transporting tools. Consider-
able effort has been made to obtain design information; however,
much remains to be done in terms of developing design equations.
Only Ashby (equation 141) has tried to develop a design equation
that will permit quantitative designs to be effected, and this he did
in 1931. All other studies have related shape to design only in a
qualitative way. Admittedly, qualitative information is needed. On
the other hand, striving to develop design equations can provide
the same information and also lead to the equations. The goals are
clear and the path is direct so that only the work remains to be done.
5.4 Manner of Movement
Once the shape of a soil manipulating tool has been selected, only
the manner in which the tool is moved through soil can be altered to
effect different manipulations. Manner of movement involves orien-
tation of the tool (angle of approach), its path through soil (depth
of cut), and its speed along the path. Shape of a tool is usually
determined when the tool is manufactured. Once fabricated, the
shape is usually not easily altered, so that the manner of movement
is generally changed during actual use in order to effect the best
performance ( 122 ). Shape is primarily controlled by the designer,
who is associated with the manufacturer. Manner of movement, on
the other hand, is primarily controlled by the user. But each tool has
inherent limitations in its practical method of operation that limit
variation in manner of movement.
The problems encountered when considering the manner of move-
ment in design are essentially the same as those encountered when
considering shape. First, the manner of movement must be de-
scribed ; second, the description must be quantitatively related to the
dependent functional factors—forces and results of soil manipulation.
Just as shape is meaningless without specifying manner of move-
ment, so manner of movement is meaningless without specifying
shape. Neither can be considered to the exclusion of the other. In-
deed, this is the crux of the design problem. Each abstract design
factor influences the operational performance of the tool. Only
256 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
CLEARANCE ANGLE
(SPECIFY PLANE) LIFT ANGLE
/APPROACH ANGLE\
\ VERTICAL PLANE/
DIRECTION OF
TRAVEL
SIDE ANGLE
APPROACH ANGLE \
.)
^HORIZONTAL PLANE/
TILT ANGLE
SLANT ANGLE \
RPENDICULAR PLANE/
10 15 20 25 30 35 40 45 50
ANGLE OF TRAVEL (•)
CO
Û-
SANDY LOAM
^ 40
CO
CO
UJ
Of 30
<
i 20H
O
2
SAND
UJ CLAY
10
LÜ
tion, which determines the lift angle, can be used to increase or de-
crease suction. This principle was used to control depth of walking
plows. By altering the hitch point, the orientation of the plow was
changed so that the downward force of suction and the upward com-
ponent of the line of pull came into equilibrium, and the plow
"floated" at the depth that produced equilibrium. Even m some
mounted plows today, the principle is used. .-..-, .. m
Tanquary and Clyde {m) designed a special hydraulically
actuated plow hitch that increases the approach angle of a tool as
the tool enters the surface of the soil. The approach angle decreases
as the plow penetrates soil and thereby decreases suction while the
vertical component of the line of pull simultaneously increases.
When the forces reach equilibrium, the plow floats and maintains
the equilibrium depth. ^ .T, •
Often the linkage system used to position a tool aflects both orien-
tation and depth. Three commonly used linkages are radial, trape-
zoidal, and parallelogram. In a parallelogram linkage, orientation
is essentially independent of depth, whereas in a radial linkage
orientation changes as depth is changed. The trapezoidal Imkage
system has intermediate characteristics. When penetration is difti-
cult, a linkage system that does not permit a tool to quickly seek
its operating depth may leave large areas at edges of fields that are
tilled too shallow. Sineokov ( 388 ) calculated roughly the length ot
path required for radial and parallelogram linkage hitch systems to
reach operating depth. While studying simple chisels and sweeps,
he secured fair agreement between the theoretical and experimentally
determined path lengths obtained for the trapezoidal linkage system.
The path required for a 27-centimeter sweep to reach operating depth
decreased significantly as the suction increased. Sineokov suggested
the following equation to express penetration relations observed for
the sweep :
S = acot(^), (142)
a = 0* a = 45** a = 90*
FIGURE 185.—Approach angle as influenced by depth adjustment with a radial
linkage system.
24 HOUSTON CLAY
20 ^-SHARKEY CLAY
16 OKTIBBEHA CLAY
VAIDEN CLAY
Û
SHARKEY LOAM
I
Ü
8
ÜJ NORFOLK SAND
OL
if)
DEPTH (in)
P.s.i.
8 2.0
10 2.1
12 2.1
14 2.1
16 2.2
1 Depth of cut wa.s constant.
SOURCE: Randolph and Reed ( SU ).
a moldboard plow that were used to vary width of cut. Shape was
inadvertently varied to the extent that different finite areas of the
same shape were used to vary width of cut. Thus, there may be an
unavoidable interaction between sliape and manner of movement in
the results.
Table 28 shows that as width of cut increased, clod size also in-
creased. The data also show that specific draft generally tended to
decrease as size of cut increased. Since lower specific draft indicates
a lower force per unit of cross-sectional area, the increase in width of
cut would appear to be advantageous. On the other hand, if a high
degree of pulverization is desired, the smaller width of cut is ad-
vantageous. The separate yet simultaneous nature of equations 1.32
and 13,3 is revealed in the results of this study.
Gill and McCreery proposed a way to consider the two contradic-
tory trends by calculating the energy api)lied to the soil for each
width of cut. They calculated the energy from draft and speed
SOIL DYNAMICS IN TILLAGE AND TRACTION 263
TABLE 2^,—Effect of size of cut on the utilization of energy in
pulveHzing soil
Equivalent
Clod Energy energy
mean- Specific Size of applied to required to Efficiency
weight draft cut soil by tool cause soil of tool
diameter (WJ breakup
(inches)
5:^300-
^ 200
H
O
UJ
100
20 40 60 80
CLAY CONTENT {%)
1000
<
Q
800
600 —1—
.8 1.0 1.2 1.4
SPEED (M/sec)
FIGURE 189.—Influence of speed on the effectiveness of electro-osmosis in re-
ducing plow draft. Solid lines, Draft values without electro-osmosis ; dotted
lines, draft values with electro-osmosis. ( Shirokov, Zhur. Tekh. Fiz. ( S85 ). )
per square inch, friction was more effectively reduced with electro-
osmosis. Dano concluded that the volumetric moisture content of the
soil was ih^ measurement that most closely indicated the effectiveness
of electro-osmosis.
Mackson ( 286 ) used a simple slider to investigate electro-osmosis
through voltage ranges of 50 to 300 volts and speeds of 2.5 to 300 feet
per minute. He determined an equation for a clay loam soil that
related the influence of various factors on the draft of the slider.
The equation had the form
D = 2.83 + OmUS - 0.00708F - OMIM + 0.514P, (147)
where D draft of slider in pounds,
S speed in feet per minute,
V potential in volts,
M soil moisture content in percent,
P normal force in pounds.
Several conclusions can be drawn from the various studies of
electro-osmosis. Its effectiveness in reducing draft seems to hinge
on the following conditions :
1. Sufficient electrical potential must be applied to cause water
movement.
2. The soil must be permeable enough for rapid water movement.
3. Sliding must occur at the soil-tool interface.
4. Good contact must exist through the soil.
5. Sufficient time must be allowed for w\ater to move.
6. The ratio of friction to total draft force must be large so that
the ehmmation of friction would be significant.
These conditions are sufficiently restrictive so that the application
of electro-osmosis will have to be specialized rather than general.
Available data indicate, however, that there are large areas in the
world where the principle may be expected to work. Nearly all
studies have been concentrated on reducing draft; but, an indirect
^ctor that must also be considered is reduced traction requirements.
Even when draft can be reduced only with an increase in total power,
traction requirements may make electro-osmosis practical or econom-
SOIL DYNAMICS IN TILLAGE AND 'ÏRACTION 269
(A) (B)
B
HOE SLICER PICK
FIGURE 192.—Typical shapes of rotary tiUer tines. (Adams and Furlong, Agr.
Engin. {!).)
quirements, in terms of both rotary and draft energy, for each shape
as well as qualitative results of soil manipulation. Thus, they con-
sidered both design equations even though sometimes they used
qualitative characterizations. A summary of their findings is shown
m figure 193. Of the three tine shapes, the so-called hoe seems to
be superior in nearly every consideration. The dotted lines m figure
193 represent the range of power requirements for small changes ot
the macroshape with no change in the basic configuration Many
workers have established the apparent superiority of the hoe tme.
272 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
ROTOR HORSEPOWER
4 -
3
2
I
HOE SLICER
I
PICK
TRACTOR HORSEPOWER
fl! H H
-I
400
I
AVERAGE HORIZONTAL 200
i1
FORCE (FORWARD)
AVERAGE VERTICAL
FORCE (UPWARD)
-200
"400
-600
-800
I I
4
1
PULVERIZATION
(AVERAGE CLOD SIZE) 2
2.0 KgM 3.0 KgM 3.0 KgM 3.8 KgM 4.2Kg^fl 5.2 KgM 6.5 KgM
FiGUKE 194.—The effect of the geometry of cut on the average torque required
to move a hoe-shaped tine through soil. (Soehne, Grundlagen der Land-
technik ( 399 ).)
from the soil mass is very nearly the same size for each soil-tool
boundary condition. The proportion of the total area that is cut by
the hoe and the proportion of the total that is sheared from the soil
mass differs for each condition. Condition A, for example, has no
shear and it has the least area requiring cutting of all the conhgura-
tions Condition F, on the other hand, has no shear but it does have
the maximum area requiring cutting. Condition B has the same
area requiring cutting as condition A in addition to maximum shear.
Condition G, on the other hand, has both maximum cutting and
maximum shear. A comparison of the average torques m figure 194
indicates that shearing requires less energy than cutting. Cuttmg
is constant in conditions C, D, and E; but shearing increases progres-
sively from zero in condition 0 to the maximum possible in condition
E and the average torque decreases as shearing area decreases, ihe
same conclusions may be reached by comparing conditions A with
B and F with G. Furthermore, as the ratio of shearing area to
cutting area decreases, the etfect of shearing on the total required
torque decreases. Even though the lateral shearing area in condition
B is much greater than the cutting area, data m figure 194 show that
torque values in A are approximately only 67 percent of those m B.
The cutting area is greater than the shearing area in condition O-
yet the torque in condition F, which has no shear, is approximately
80 percent of that in G. A comparison of conditions B and £", where
the ratio of shearing area to cutting area is nearly equal, shows that
the torque in B is approximately 71 percent of that in A. it the
cutting and shearing components of the total torque were equal, con-
dition B should require more than twice the torque ot condition A.
Similarly, shearing should have a greater influence m the other con-
ditions if shearing and cutting were equal. Thus, shearing soil
appears to require less energy than cutting. .. . A +;„oa
Soehne also studied macroshape and orientation of hoe-shaped tmes
with the same experimental apparatus. He expressed measured
torque values on the basis of the volume of disturbed soil (size ot the
piece) in order to evaluate the energy requirements on a specific work
basis. He calculated the specific work by dividing the average torque
by the volume of disturbed soil. Figure 195 shows some of the tool
parameters that he studied. He did not describe the exact shape ot
each tool; but the radius R and width W are parameters that reflect
274 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
CUTTING ANGLE
O O
he passed through the soil once. The rake angle (fig. 195, B) has
no significant eíTect on specific work over a range from 50° to 80 ,
whereas the cutting angle (fig. 195, C) seems to have considerable
effect Data in table 81 indicate that minimum energy is required
at approximately 20°. The cutting angle 8 (fig. 195, G)_ cannot
be considered by itself since the clearance between the tine and
the soil is also present.
Figure 196 shows the relation between the clearance angle a and
the cutting angle 8 for a rotating tool. The tine angle of sharpness
/
/
I
/
\
\
\
\
N
B
FIGURE 19G—A, Relation between clearance angle a and the cutting angle 8 of
a rotary tine; B, cutting and clearance angle correction associated with
forward implement velocity u,.
path, the relation between the actual cutting angle and the angle
correction factor A8 ranges from zero where the implement and tool
velocities are parallel to a maximum where they are perpendicular.
The relation shown in figure 196 can be used in either graphical or
analytical methods to determine true cutting and clearance angles
for any combination of speeds and depths of operation.
Umax'ZB6Mm
1+2'
r+2 r+2
1 + 2'
90» 180«
FIGURE m.-Effect of number and location of tmes «", tofX " 2 HnP roíor
rotor with tines operating as pairs in tlie same row ; bottom, 12-tine rotor
with Tines equally spaced around the rotor. (Soehne, Grundlagen der
Landtechnili (399).)
for one tine 31 and two tines 231 passing through soil. For a rotor
of the usual diameter operating at the usual depth, all soil resistance
is encountered in a 90°-angle of rotation. As figure 197 shows,
peak torque 31„a. and average torque J/,„ for an implement can diner
greatly because of the arrangement of the tmes about the rotor.
Proper spacing can decrease the variation in amplitude of torque to
provide a smooth and desirable torque pattern. Increasing the
number of evenly spaced tines on a rotor produces a snioother
torque curve. Also, with more tines on a rotor, the peripheral speed
can be reduced while, the size of the soil slice is maintained tor a
given forward speed. On the other hand, spacing tmes too closely
causes clogging, and slower rotating speeds decrease cutting and
clearance angles. Any final arrangement and spacing of tmes, there-
fore, becomes a compromise. . í . ^;n<,^
Figure 198 shows the effect of three arrangements of rotary tiller
tines on cut and torque patterns. The technique shown m figure 198
is one means of evaluating the geometry of cut for any particular
tine arrangement. Careful examination of figure 198 shows that
individual tines can have different boundary conditions, depending
on the arrangement. For example, in some tine arrangements, the
278 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
180°-,
^ 90
T 1 r-
12 3 4 5
,V'/v\^^^vV^v\
actual size of the soil slice differs for different tines. Even if the
size of the slice is constant, the nature of soil failure along the
boundary may vary. Detailed consideration of soil-tool geometries
shown m figure 198 shows that the cutting and shearing areas of a
soil slice vary from tine to tine. Unless the rows of tines run pa-
rallel to the rotor axis, the cutting and shearing areas for individual
tmes cannot be made to be identical. Interior tines shown in the
flat spiral arrangement at the left of figure 198, for example, have
shearing areas of soil slices that alternate between 40 and 60 percent
of the length of cut—that is, the shearing area of an external tine.
Ine steep spiral arrangement at the center has a shearing area that
alternates between 20 and 80 percent of the length of cut. The
unbalanced arrangement at the right, which lets the two external
tmes strike at the same time, should be avoided. In spite of the
variation m shearing area for the steep spiral, its arrangement is a
suitable compromise of all factors.
In addition to the gross arrangement of tines around and along
the length of a rotor, the "microarrangement" of tines is important.
± igure 199 illustrates how the microarrangement of individual tools
may be varied to control overlapping and to improve performance
Arrangement C requires the least power; arrangement A, the
greatest. A careful examination of the geometry of cuts in figure
199 shows that arrangement G tends to minimize the drag between
an individual tme and the soil slice that is cut loose. The frictional
drag IS minimized because the overlapped tine moves along an area
where the soil slice has already been largely formed by the passage
SOIL DYNAMICS IN TILLAGE AND TRACTION 279
< 3
er.
<
3
O
10 15 20 25 30 35
TOTAL FORCE (lbs)
FREQUENCY (CPM)
thrust of the tool (fig. 202, B) will have to slide along the surface
of the tool. Gunn and Tramontini ( 167 ) have clearly demonstrated
that the total energy (draft energy plus oscillating energy) can
be reduced. Figure 203 shows the total energy requirements for a
rigid tool and the same tool when oscillated. Even when total
12 3 4 5 6
TRACTOR VELOCITY-FT/SEC.
FIGURE 203.—Total horsepower requirements for a tillage tool when rigid and
when oscillating. (Gunn and Tramontini, Agr. Engin. {161).)
PATH OF
CUTTING TIP
A B
FIGURE 204.—A, The direction of oscillation of an inclined oscillating tillage
tool; B, the path of the cutting tip. (After Eggenmüller, Grundlagen der
Landtechnik {110).)
the path of the tool is described by the arc of a circle. For small
amplitudes, the arc can be considered to approximate a straight
line. Figure 204, B shows the actual path the tip of the tool makes
when amplitude, direction of oscillation, and forward speed are
fixed.
Eggenmüller developed mathematical equations that describe the
path of motion where the oscillating motion lies in a straight line
m the implement reference system. The position of a chosen point
in the implement reference system, such as the tip of the tool,
could be easily described at any time t if the oscillation is sinu-
soidal. The position of the tip at an}^ time t lies on the path of
motion BD (fig. 205). The line of oscillation he is oriented in the
implement reference system at the angle j) with the horizontal, and
the position of the tip at any time t is given by the relationship A
sin Cut where A^ o), and t are, respectively, amplitude, angular fre-
quency of oscillation, and time. The implement reference system
moves with velocity v in the earth reference system during oscilla-
tion; when the tip is at point a in the implement system, it is
simultaneously at point B in the earth system. Disregarding oscilla-
tions, point a will move to point C in the earth system at time ^,
and the distance B — 0 \& equal to v t. During time t the tip will
have moved from a to c m the implement system, but point a in the
implement reference system will have moved from B to C in the
earth system. The combined motions thus result in the tip moving
from B to D m the earth system. The distance B — D is the vector
284 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
Y
IMPLEMENT REFERENCE
SYSTEM
EARTH REFERENCE
SYSTEM
6 9 12 .4 .6 .8
AMPLITUDE (mm) DIRECTION OF SPEED Vo(m/8)
OSCILLATION (*»)
RIGID
^ 800 O
o
- 600
OSCILLATED
< 400
(r
Q
200
(B)
FIGURE 208.—A, A double-cut plow ; B, the tillage action of a double-cut plow.
(Domsch, V.E.B. Deutscher Landwirtschaftsverlag, Berlin {lOJ^).)
SOIL DYNAMICS IN TILLAGE AND TRACTION 291
208 indicates the principle of a double-cut plow that was designed
to break up a dense layer below the normal plowing depth. The
double-cut plow is also useful when infertile or wet material should
be kept separate from the surface soil. Figure 208 B shows how the
lower share may loosen and pulverize the soil but leave it in the
furrow bottom without any appreciable mixing with the surface
layer. Notice that the lower cut moves soil upward into an uncon-
fined area. As a result, the same volume of soil can be tilled in
layers with less energy than is required to till it in a single cut.
Table 34 shows how draft force is reduced by using the double-cut
principle {357), The same volume of soil was tilled by the two
types of plows; however, the double-cut plow required 25 to 30 per-
cent less draft force than the conventional plow. Notice also that
the various depth settings of the double-cut plow produced different
total drafts for the combination. While the interaction appears to
be small, the effect can be truly evaluated only when each portion of
the double-cut plow is measured simultaneously.
(A)
(B)
(C)
k<H kb-l
(A) (B)
The spacing and width of teeth along the cutting blade were also
investigated by Zelenin. In a series of experimental arrangements,
the ratio h/a shown in figure 211, B was varied. The results are
shown in figure 212. The data indicate that a definite minimum
occurs at a h/a ratio of approximately 2.5. The curve is based on a
limited number of points, however, so that the exact magnitude of
values may be questioned. Since a ratio of zero indicates an in-
finite number of teeth and a ratio of infinity indicates no teeth, the
data do not appear too unreasonable in light of the data included
in table 35. An interaction between teeth on a cutting blade is
clearly demonstrated by these data.
Length of teeth on cutting blades was also investigated by Zelenin.
For a given orientation (angle of cutting a in fig. 211, Ä), an in-
crease in tooth length results in a greater clearance distance X below
the cutting blade. Added length was detrimental because it de-
creased the strength of the tooth, increased resistance to cutting, and
increased "littering" (some of the loosened soil slipped between the
SOIL DYNAMICS IN TILLAGE AND TRACTION 295
1.0 ■
X X
h- K
UJ U
UJ LU
1- h-
X O
§ 1 0.8
UJ LÜ
(y fy
Q Q
ü- U.
O O
2 Z
h- H
h-
Z5
t
3
Ü O 0.6
teeth and was not placed in the scoop). Based on these premises
an equation of the form
(151)
X = L sin
{y^i}
where X — clearance distance,
L — tooth length,
y — cutting clearance,
ß = angle of sharpness of tooth,
expressed relations that affected a suitable design. Based on ex-
perience, Zelenin reported that the optimum cutting angle a should
be about 20° for most conditions. A maximum angle of sharpness
is thereby specified, since a minimum cutting clearance angle of 5°
is considered essential. An angle of cutting of 30° was considered
to be maximum, so that a compromise between strength (large
angle of sharpness) and lowest draft (20° angle of cutting) has to
be effected.
A final point may be made about tillage implements. Figure 213
shows an example of a drawn tool whose action is modified by the
dynamic action of a power-driven rotor. Endless possibilities exist
for developing multipowered implements, particularly since engine
power ceases to be a limiting factor. Interactions between complex
tool systems of this type are essentially unexplored.
5.7 Principles of Force Application
To design, by definition, is to conceive a scheme or plan. Since
design implies intent, it becomes important to judiciously choose
296 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
and vary design factors in order to obtain the best possible design.
Two tillage equations described in section 5.2 can be used in develop-
ing optimum designs or in determining the influence of various de-
sign factors on performance. Nothing about the factors to be
selected can be determined from these equations. As a result, it
becomes necessary to choose design factors which, by experience or
intuition, seem to be of basic importance and study them. It is
within the scope of soil dynamics research to provide information
that will ultimately assist in identifying and selecting design factors.
At present (1965), little information is available; however, the
general nature of information that will be required can be foreseen.
To have the skill necessary to design tillage tools that will operate
in the most efficient and most effective manner, the basic principles
of the tillage actions must be identified, developed into distinct
concepts, and properly used. These principles should be used as a
guide to direct the application of forces and movements of tools
in order to achieve the desired soil conditions. Following are illus-
trations of several basic principles of force application:
1. Forces should be applied to the soil in the direction that re-
quires the minimum force to cause the desired action. When only
a cutting action is desired a flat blade can be used to cut the soil
without applying force to lift and loosen it.
2. The mechanical rigidity of the soil mass should be used as a
holding body. The application of force in the desired location and
SOIL DYNAMICS IN TILLAGE AND TRACTION 297
direction can be controlled when the soil is rigidly held. The direct
fragmentation to the desired clod size may be better achieved by
using the inherent strength of the undisturbed soil mass as a holding
body than by disrupting the mass and then reducing the size of
granules by repeated operations. .
3. The soil-tool geometry should be designed so that the confining
action of the soil is either minimized or utilized in conjunction with
the direction of force application of the tool. Small cuts may reduce
the total force on a tool due to reduced soil confinement or due to the
lesser weight of soil on the tool. The removal of soil from the
cutting areas of augers, ditchers, etc., prevents binding and the need
to displace soil against a high mechanical resistance. The direction
of force may be controlled to utilize the confining action of the
soil mass. In packing or crushing operations with rollers, a large-
diameter roller traps surface clods more successfully than a small-
diameter roller, which tends to roll them in front of the roller.
4. The rate at which force is applied should be controlled. Im-
pact forces may be applied to detached soil fragments in order to
exploit soil inertia and build up stresses in the fragments. Brittle
clods may be broken without entrapment and confinement by the
soil mass.
5. The sphere of influence of an application of force should be
controlled. Changes in strength caused by strain hardening result-
ing from soil compaction may be undesirable or they may influence
the reaction of the soil to subsequent tillage actions. Extraneous
forces such as those imposed by secondary forces—that is, subse-
quent machine operations—frequently undo the work done on the
soil by primary tillage forces. The size and shape of tools also
govern the sphere of influence of the tool.
6. The force applications should be directed to control the size
and shape of detached soil blocks. The shape and path of move-
ment of the tool may be used to influence the size and possibly the
stability of soil blocks.
7. Balanced force applications should be used for multiple tool
units. The controlled movement of several tools can insure or-
der and regularity in the action of the tool on the soil and result
in a more uniform and efficient use of the power input.
8. A sequence of forces should be applied to execute complex ac-
tions. The use of a coulter preceding a plow cuts plant material and
establishes order in a randomly spread plant material so that effec-
tive covering can be accomplished by the plow.
9. The time of force applications in a system should be controlled.
Breaking clods immediately after plowing will preclude the develop-
ment of high strength by drying before breakdown is effected.
Waiting for drying, on the other hand, increases strength so that
compaction is resisted.
There should be a continued search for other principles since they
provide a sound basis not only for the design of tillage tools but
also for the subsequent development of methods by which they
should be used. The development and use of principles for design
will permit a more direct and accurate selection of design factors for
study in the design equations. Unless this is done, pertinent factors
may not be introduced as quickly as possible.
6. PERFORMANCE OF TILLAGE TOOLS
6.1 Introduction
Design was discussed in chapter 5 without regard to use of tillage
tools. Disregarding application was convenient because it simplified
the study of design. Application cannot always be ignored, be-
cause the ultimate purpose of design is not to build tools but to
change undesired soil conditions into new soil conditions that will
better serve some specific intended use for the soil.
Performance is defined by Webster as the act of performing. To
perform is to accomplish, or to carry on to the finish. Performing
implies action; and performance can have almost as many specific
meanings as there are specific actions. In tillage, the obvious action
is the manipulation of the soil into a different condition. Perform-
ance of a tillage tool thus may be defined as the production of a
change in soil conditions by manipulation of the soil. Performance
includes two distinct and separate factors of interest : the amount of
soil manipulation, and the magnitude of forces required to cause
the manipulation. These factors must be quantitatively assessed in
order to measure tillage tool performance.
Each factor of tillage tool performance is fixed, since it does not
vary for a fixed manipulation. As indicated by tillage equations
132 and 133, in chapter 5, the use of a given tool in a given soil
condition results in a fixed change in soil condition. The difference
between the initial soil condition Si and the final soil condition Sf is
a measure of soil manipulation. Similarly, the forces required to
cause the manipulation do not vary for a fixed manipulation. Given
the necessary information and the idealized type of complete mech-
anics discussed in chapter 4, tillage performance could be calculated.
Similarly, if the tillage equations were fully developed, they could
be used to calculate performance. Tillage performance, when de-
fined in this sense, is a fixed and determinable entity that can be
used to express the performance of a soil-tool system.
Performance embraces more than just changing the soil to a new
condition. First, the tillage tool should perform efficiently ; that is,
soil conditions should be altered with the smallest expenditure of
energy required to operate the tool. Second, the final soil condi-
tion must be acceptable. A tillage tool may manipulate soil very
efficiently, but the soil condition it produces may not be suitable
for the intended use of the soil. Performance, in terms of suitability
of soil conditions, must be evaluated by comparing the final soil
conditions that are produced with those that are desired. Perform-
ance, in terms of forces, must be evaluated by comparing forces
required for manipulation with available or acceptable forces.
The soil conditions that are to be produced by tillage must be
expressed in criteria that characterize the soil in terms associated
298
SOIL DYNAMICS IN TILLAGE AND TRACTION 299
retical but are separate and distinct. That the reactions can and
often do occur simultaneously does not negate the principle of their
independence; it merely complicates the situation. The important
point, however, is that any use of soil involves either an active or a
passive behavior of the soil. These behaviors can be represented by
suitable behavior equations. As was discussed in section 3.1, the
equations contain parameters and the parameters provide a means
for numerically describing the role of the soil in the behavior.
Since the parameters are defined by a behavior equation that repre-
sents the specific behavior associated with some intended use, be-
havior parameters provide the mechanism by which to describe soil
conditions in terms that are compatible with the intended use.
Examples of active behavior equations and various means for
assessing the parameters were given in chapters 2 and 3. Active
soil behavior is relatively easy to visualize because the soil itself
actively participates in the behavior. This active participation is
the movement and yield of the soil. Passive soil behavior is dis-
tinctly different from active behavior. In spite of this distinct dif-
ference, the parameters of the equations that represent the respective
behaviors serve the same function, namely, they assess the role
the soil plays in each respective behavior. j, r\^ j
To visualize passive behavior, consider the application of Ohms
Law to soil. In this application, the soil acts as a "conduit"^ or a
"conductor" for the now of electrons; it serves as a transmitting
medium. During this behavior, the soil itself is not active ; that is,
it is not moving. Eather, by its rigidity it is guiding and con-
trolling the active material, the electrons. Also when air or water
moves through soil, the soil without motion guides or controls the
flow of the active material. Thus, in passive soil behavior the
transmitted material is the active material, whereas in active soil
behavior the soil itself is the active material. Furthermore, passive
behavior of soil always involves two materials—the soil and the
material being transmitted. These two materials may be envisioned
to be a combined system, and an overall behavior equation will
represent the behavior of the system. This same type of relation
applies to the soil-tool system where the soil is the active material
and the tool guides and controls the flow of soil. The functional
relation provided by equation 132 (discussed in ch. 5) thus might
serve as a passive behavior equation of a tillage tool.
It is important to separate soil behavior into active and passive
reactions. Theoretically no conflict exists, since the separation is a
mental model of the physical behavior; in practice conflict does
exist, however, because active and passive reactions can and do
occur simultaneously. Furthermore, force systems can and do
vary continuously in magnitude. Whether the same force system
causes the soil to enter into a passive or an active reaction depends
on the magnitude of the forces. In other words, there is no way to
look at forces and separate them into those that will cause passive
behavior and those that will cause active behavior. In practice,
however, a separation can be made. To illustrate, consider water
flowing through a pipe. Characteristics of the pipe such as size,
shape, length, and roughness guide and control the flow of the water
through the pipe—that is, they determine the role of the pipe m the
302 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
. i/=
v = V4=5:(v,.v,.av8)
Vsb
X » VALUE OF COUPLE = Va
T(V| ±V2)
(A)
"%flT
FIGURE 215.—A dynamometer employing .«train gages for measuring soil forces
on tillage tools.
SOIL DYNAMICS IN TILLAGE AND TRACTION 313
TRAVEL
iC
UJ
(T
(/)
(/)
UJ
cn
a.
ÜJ
o
¡2
01
DISTANCE (in)
relation of pressure near the tip and the draft. The cyclic nature
of the curves coincided with the development of major shear failures
m the soil. Mauer also investigated the relation between the lift
angle of the chisel and the maximum measured unit pressure as
affected by soil preparation factors, moisture content, and the pres-
sure used to compact the soil.
The data in figure 218 show that any factor that increases the draft
of the tool also increases the pressure on the surface of the tool.
Instrumentation techniques have improved considerably since Mauer's
work, but even recent studies ( 332, ^19 ) have not produced detailed
information about the distribution of pressure on the surface of a
chisel.
Mayauskas ( 295 ) has determined the pressure distribution on the
surface of a plowshare. A series of pressure transducers were placed
m a pattern so that simultaneous measurements could be made (fig.
218), and a series of experiments were conducted in which the instru-
SOIL DYNAMICS IN TILLAGE AND TRACTION 315
PRESSURE OF
10 SOIL COMPACTION
10 /"^ 7psi
Q. 8 - 3 8
UJ
LÜ a: 5psi
£r
(/) 6 (/) 6
UJ
UJ a:
Q_
Q.
2 4 3psi
S 4
3
X
< <
2 2
S 2
15 25 35 45 18 22 26 30
mented plowshare was used. The results (fig. 220) show the man-
ner in which the factors investigated inñüence the magnitude and
distribution of pressure on the surface of a share. No information
is available to indicate what magnitude or distribution ot pressure
can be considered to be good performance. Certainly, however,
criteria could be developed and performance measured m terms ot
the distribution and magnitude of pressure on the soil-engagmg sur-
face of tillage tools. Measurements of this type have been made
in several countries. . .
In another approach directed toward measuring performance, an
attempt was made to separate the draft of a moldboard plow into
4 4 -^ 4
■
-^
-ß
\
3 3 -^ 3 I\\
2 2 2 - vvsii^^y^
1 1 I P^^^o o 1
—1—1—1—1—1—1—I 1 1
1.0 2.0 3.0 4.0 1.0 ^0 3.0 4.0 1.0 2.0 3.(
PRESSURE (Kg/cm*)
(A) (B) (C)
portions required to cut, pulverize, and invert soil. This was done
by cutting a plow into sections and measuring the draft of the sec-
tions as they were progressively added together (table 37). The data
indicate that the final one-fourth of the moldboard did not materially
add to the draft at either of the two speeds studied. This final sec-
tion, however, may have been of paramount importance in obtaining
adequate inversion of the soil. Pounds of draft per se, therefore,
cannot be used as the sole criterion for performance. As was the
case tor surface pressure distributions, no standards are available to
indicate whether the performance indicated in table 37 is desirable
At the very least, however, the data are useful for a qualitative
understanding of the interaction of tool and soil.
A tool inclined 3°, together with a wire, has also been used in an
attempt to separate cutting, sliding, and shearing actions for a simple
tool (352), The draft force required to pull a wire through soil,
as shown in figure 77, may be interpreted as being the force required
to cut or separate the soil slice from the semi-mfinite undisturbed
mass of soil. The thin, ñat blade operated at a lift angle ot 3
separates a mass equal to the mass of soil separated by the wire.
The total force both cuts the soil and overcomes the friction ot the
soil sliding over the upper surface of the blade. The 3° lift angle
provides clearance so that friction does not occur on the lower sur-
face of the blade. . i p^ j! ^i.
The difference between the draft of the wire and the dratt ot the
inclined blade was assumed to be due to soil-metal friction. It the
blade is operated at a lift angle of 221/2° or 45% the ridge of soil will
not only shear but also pulverize. The draft of the blade may be
thought to be composed of forces required for cutting, sliding (fric-
tion), shearing, and accelerating the soil. on-. j
The difference between the drafts of the wire, the 3 blade, and a
blade with a higher lift angle permits separation of the draft com-
ponents required for cutting, sliding, shearing, and accelerating the
soil. Figure 221 shows the measured forces as a function ot speed
500
400
300
<
a:
Q
200
100
SPEED (mph)
FIGURE 221.—Effect of speed on the draft of a wire and a flat inclined plate.
as the wire and blade were operated in a clay soil. These forces may
be separated into the relative contributions of cutting, sliding fric-
tion and shear plus acceleration. Visual observation indicated that
the amount of soil breakup was nearly the same when the blade was
inclined 22i/4° and 45°. Figure 222 shows the contributions ot the
various actions when the draft of the 221/2° blade was assumed to
represent the total required draft. It appears that speed greatly in-
fluenced the relative contributions of the origins of resistance, ihe
high percentage of draft that can be attributed to cutting is also
significant. Cutting represented a large component of the dratt ot
the moldboard plow reported in table 37.
318 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
CUTTING
SHEAR + ACCELERATING
SPEED (mph)
FIGURE 223.—Rotary sieve for separating fragmented soil into various sizes.
3r
< 2
o
Z5
UJ
<
a:
ÜJ
12 3 4 5 6 7
DEPTH OF CUT (in)
resulted when the depth of a rotary tiller was varied. The data
appear to contradict those in table 38, since clods were smaller at
deeper depths of operation. Because the movement of drawn and
rotary tools differs, the data may not be in conflict. The depth of
operation of a rotary tool may be changed without greatly altering
the size of cut. A change in either forward speed or rotor speed
materially changes the size of cut. Furthermore, a vertical gradient
usually exists in moisture and strength in a soil profile. Soil at the
deeper depths may fragment easier so that smaller clods result. In
SOIL DYNAMICS IN TILLAGE AND TRACTION 321
TABLE 38.—Effect of the size of cut on the mean weight diameter
of clods produced in a compacted clay soil
Size of cut Clod size in mean
Tool by tool weight diameter
Inches Inches
Plow and coulter 1 1.47
2 3.55
4 6.46
6 7.07
8 8.61
Disk plow 1 1.76
2 3.66
4 4.49
6 4.19
8 4.15
1 SOURCE : Gill and McCreery {IJ^d ).
4r
tr 3
LÜ
<
Û
UJ
û:
ÜJ
FIGURE 225.—Effect of increased impact energy on clod size for various sizes
of cut of a rotary tiUer. (Adams and Furlong, Agr. Engin, (i).)
322 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
but rotor speed and forward speed must both be carefully controlled
with rotary tools. By increasing the forward speed in accordance
with the rotor speed, a uniform size of cut can be maintained with a
rotary tiller. Figure 225 shows the results oï an experiment where
the depth of operation was 4 inches and several sizes of cut were
used. As the rotor speed increased, the cutting speed of individual
tmes increased so that they passed through the soil in less time.
Higher speeds produced greater impacts. The data show that as
the impact increased, the average clod sizes decreased even though
the size of cut was constant. Table 39 shows .a similar trend for
drawn tools. The change in speed and, hence, change in impact is
much less for the chisels reported in table 39. The change in
breakup is also less, but the data show that increased impact causes
increased breakup.
6.4,2.2 Segregation
Nothing in section 6.4.2.1 describes the final location or position
of soil clods in the soil profile ; emphasis was on a description of the
change of clod size. Tillage is sometimes undertaken to control the
distribution of clods within the profile. For example, large clods
may be placed on the surface of a soil because they tend to prevent
wind erosion ( 510, 511 ). On the other hand, small clods are usually
desired in the vicinity of seed to provide an optimum environment
for germination. In such instances tillage tries to segregate soil
according to clod size or some other criterion that will improve the
condition of soil. Sieving distinct layers of the soil profile provides
a means for measuring the segregation performance of tillage tools.
The action of narrow tines on tillage tools such as harrows segre-
gates clod sizes. Winkelblech ( 507 ) used the sieving technique to
study the effect of several common tools on the distribution of clods
in a soil profile that was uniform before tillage. He used a pulveri-
zation modulus proposed by Yoder ( 5H ) to express clod size. The
modulus is essentially a weighted average ; the average is so weighted
that the resulting modulus reflects an average clod size in a sample.
Winkelblech used size classes li/4 to % inch, % to % inch, % to %6
inch, %g to %2 inch, and less than %2 inch. To calculate the pul-
verization modulus, the percentage of soil in each size class is multi-
plied by an ordered weighting factor. Winkelblech used zero for
less than %2 inch, 1 for %6 to %2 inch, and on up to 5 for the
SOIL DYNAMICS IN TILLAGE AND TRACTION 323
largest size. The weighted percentages were summed and divided
by 100 to give the pulverization modulus. Thus, if all the clods were
in the largest size, the calculated modulus would be 5. On the other
hand, if all the clods were less than %2 inch, the modulus would be
zero.
Winkelblech determined the pulverization modulus for 1-inch
increments of depth in an experimentally prepared soil after it was
tilled to a depth of 4 inches. The untilled soil had a uniform modu-
lus at all depths. Figure 22^6 shows the segregation of clod sizes that
PULVERIZATION MODULUS
SPRING TOOTH
z 1-2-
DISK 18*»
DISK 28**
Q. SWEEPS, 10"
ÜJ
O
2-3
-UNTILLED
COARSER
xxxxxxxxxxxxxxxxxxxxxxxxxxxxx
FIGURE 227.—Segregation of subsoil to the surface to permit soil renovation.
CLOD
ARRANGEMENT
AUGERS
LIFTING BLADE
^SIEVE SLOTS
seedbed while large clods are placed between rows where control of
erosion is important. The device utilizes small clods that are already
present to form a seedbed so that additional energy is not required
to further break soil into small clods. The principle of segregation
of clod sizes may save more work in tillage than any other principle
that has been proposed. The positive control permitted by sieving
may make the principle practical. Sieving thus is not only a means
for measuring segregation performance ; it is also a means for effect-
ing the performance.
6,4,2,3 Mixing
Often an objective of tillage is to mix the soil to obtain uniform
distribution of moisture or clods—that is, a nonsegregated soil con-
dition. In other instances, granulated material similar to soil ag-
gregates, such as fertilizer, needs to be mixed with soil. Uniformity
of mixing is a measure of mixing performance.
Tracer materials have been used to measure mixing performance.
Wooten, McWhorter, and Eanney (512) used fluorescent materials
that could be photographed to obtain a graphic representation of the
degree of mixing. These workers reported that at least three disk-
ings are required to adequately mix material that is applied to the
surface. Hulburt and Menzel ( 188 ) used grain sorghum and radio-
active phosphorous as tracer materials. The tracer material was
located by excavating to expose a vertical profile of the soil for
observation or sampling.
Figure 229 shows the results of two tillage treatments in which
grain sorghum was used as a tracer material. Eotary tilling should
be done at least twice, according to Hulburt and Menzel, to obtain
adequate mixing. Tracer materials provide a convenient means for
measuring mixing performance.
GROUND
c I -.
¿SURFACE
t 5-
7
K 30in •
(152)
2 {Xi-ü,y
where N - number of spot samples taken from a mix containing a
mean fraction of additive JJa^
Xi = indicated content of any sample,
CTn theoretical standard deviation at zero mixing,
V¿Va(l-¿/„),
standard deviation at any time,
{x,-üay
n
This coefficient may be useful when fertilizer or some similar gran-
ular material is to be mixed in soil. The approach can be modified
to suit other situations. The concept, however, provides a means for
expressing mixing by a single number and so provides a means to
determine mixing performance.
6.4.3 Specialized Tillage Actions
In section 6.4.2, several means were discussed of measuring tillage
perforniance where emphasis is placed on the condition of the soil.
Often tillage involves a specific action where condition of the soil is
of secondary importance. Since these actions are usually highly
specialized, they are classed here as specialized tillage actions. Ex-
amples are handling plant residue, inserting foreign materials such
as dram tile or electric cable into soil, and separating root crops such
as potatoes or sugar beets from the soil.
Since soil condition is of secondary importance, performance is
concerned with forces required to operate the specialized tillage tool
and with the completion of the intended action. Often the action
does not require a numerical description of the degree of completion;
either the action is performed, or it is not. For example, either an
electric cable is buried in a suitable operating position, or it is not.
Measuring performance of specialized tillage actions is often rela-
tively simple when compared with measuring performance where
soil condition is of interest. If desired, degrees of success might be
introduced m terms of the number of cable breaks per mile, the
number of points of inadequate depth of placement, or some other
limiting condition.
6.4.3.7 Handling Plant Residue
Tillage is often performed where plant residues are encountered.
In some instances the residue must be handled only to prevent its in-
terference with the operation of the tillage tool. Usually, however,
SOIL DYNAMICS IN TILLAGE AND TRACTION 327
the residue is to be incorporated into soil and handling requires more
than just preventing interference.
Plant residues are occasionally handled by auxiliary devices used
with a tillage tool. The most widely used attachment is probably the
coulter that is associated with a moldboard plow. The use of auxil-
iary devices is a departure from the philosophy that a single tool
can be pulled through soil to accomplish everything that is desired.
When soil conditions are to be changed, such a philosophy can prob-
ably be applied without any theoretical limitations. When special-
ized tillage actions are involved, however, theoretical as well as prac-
tical limitations suggest that a multistage action is required to
accomplish the intended change. In such instances, multiple tools
probably will best meet the requirements.
The performance of auxiliary devices has received considerable
attention. One reason is that the performance can be assessed with-
out measurements. Coulters, for example, perform a definite cutting
action that severs plant residue. Cutting changes the dimensions of
the residue so that its maximum width is the width of a furrow slice.
With the plant material cut, the natural action of lifting and turn-
ing a furrow slice by a moldboard plow is not hindered, and the resi-
due is covered by the soil. Performance of the coulter, therefore,
can be easily determined without measuring by observing how effec-
tively the material is cut.
A number of variations of coulters have been used. Knives,
notched coulters, and disks have been designed for use with mold-
board plows. Notched coulters have been powered to act as a saw,
and it has been reported that they are effective in cutting heavy
trash {390), The unpowered notched coulter traps the plant ma-
terial so that it is cut as the coulter attempts to force it into the soil.
This trapping action minimizes the bulldozing of the residues that
decreases the cutting effectiveness of a coulter. For effective cutting,
the soil conditions must be firm ; otherwise, instead of being cut the
material will be crushed into the soil. The crushing action has been
utilized with blunt tools to anchor plant residue to the surface of
soil in order to minimize wind erosion {73), Disk packers with
spaces of 4 to 6 inches were effective in pressing plant materials into
soil to control erosion.
A jointer is another auxiliary device that has been used in hand-
ling plant residue. A jointer moves a portion of the surface residue
toward the open furrow. This detached residue is literally thrown
into the bottom of the furrow and covered. A jointer also tends to
break up the soil it contacts, but its primary purpose is to aid m
handling plant residue. A combined coulter and jointer handles
plant residue effectively. Plant material that is extremely long or
that lies parallel to the direction of travel, so it is not cut by the
coulter, can often be oriented by means of a wire or trash guide.
These devices guide the material along an intended path or press
the material tightly against the soil so that it is covered by the soil
during plowing. Performance of devices for handling plant residue
usually is not measured; it is determined by the effectiveness with
which the residue is handled.
Figure 230 shows how three plowing techniques place surface resi-
AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
CGMMUNICATKDN
"^ WIRE
(A) (B)
FIGURE 232.—Soil reaction resulting from the use of a chisel operated : A, With
a small approach angle; and B, with a trailing approach angle.
110
LOAMY CLAY
100
HEAVY CLAY
90
HUMIC LOAM
^ 70
uj 60 SAND
o
g 50
11.
40
o
z 30
t 20
10
O
6 8 10 12 14 16 18 20 22 24 26
ROOT LENGTH (cm)
inclined lifter disturbs more soil around a sugar beet and therefore
may decrease breakage of the beet. If a significant portion of the
beet is left in the soil, the method becomes economically impractical.
A wide angle between the lifting blades was reported to leave enough
soil in front of the beet so that soil resistance prevented the beet
from bending and breaking during extraction. The energy require-
ments of a steeper lifter may be greater, but a more effective extrac-
tion could justify the increased expenditure of energy.
V-/7777777.
(A)
FIGURE 234.—Partial excavation and Ufting of roots. (Hülst, Gohlich, and
Sochting, Zucker ( 189 ).)
334 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
the two factors could be weighted in the combined term. This con-
cept has no limitations as to what constitutes a performance factor
other than that the factor can be interpreted in a meaningful manner.
Obviously, extreme care has to be used in devising a performance
factor so that its physical significance is clear. Performance factors
are often necessary when no clear-cut performance measure can be
made, and they can be extremely useful in developing terms that can
be evaluated.
Maximizing or optimizing a performance factor does not indicate
whether the performance is acceptable or even desirable. That de-
cision can be made only by evaluating the performance factor. The
difference between the numerical values of actual and desired per-
formance evaluates performance. If the values differ widely, the
performance may be judged to be poor even though the particular
factor may be maximum for the situation. The complete independ-
ence of performance and evaluation of that performance must be
recognized in order to fully utilize the concept.
The concept of evaluating performance by comparing desired and
actual performance is simple and direct. Unfortunately, desired
performance cannot always be expressed in meaningful, finite terms.
For example, the capacity of a tillage machine is usually desired to
be as high as possible ; only in unusual circumstances will a definite
capacity be desired. Similarly, the power requirements of a tillage
machine are usually desired to be as low as possible. Since an in-
crease in the capacity of a tool can be secured only with an increase
in power requirements, a practical balance between the two factors
must be attained.
One means of overcoming the limitation in evaluation when de-
sired performance does not have a fixed value is by using an evalua-
tion factor. The evaluation factor is an extension of the concept
of the performance factor. An evaluation factor is a term used to
express the composite effect of individual performance factors. Eval-
uation factors must be devised so that meaningful interpretations can
be made; this requirement for interpretation is the only rule that
guides or limits the development of a specific evaluation factor.
The following general guidelines can facilitate the development of
an evaluation factor. Each available fixed value of desired per-
formance should be used as a separate term in the evaluation factor.
Each term can be the desired performance divided by (the desired
performance plus the absolute value of the difference between the
desired performance and the actual performance). This technique
produces a term that has a maximum value of 1 when the desired
performance and the actual performance are equal. The term ap-
proaches a value of zero as the difference between the actual perform-
ance and the measured performance approaches infinity. If several
such terms can be devised, the product of the terms would provide
a meaningful evaluation factor. A value of 1 indicates that all
actual and desired performance factors are equal and that the de-
sired performance was attained. A value of less than 1 indicates
that the desired performance was not attained by one or more of the
individual performance factors measured. When individual per-
formance terms are not of equal importance, they can be weighted to
336 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
fe ^ J::^ r^ ^'^ (M lO o rH o (N 00 00
í^ O
Tfi Tt^ ÍO b- (N
ÇOTHTHrHC^_ OOrHrH Ci Oi 00 (M (M
Ö TA Oi (M* CO rH (TÍ (M*
fO ÇDTt^CO00t^ OOOOO 05 05 t- O 00
■^ iO 00 rH r-l CO t^ CO CO OOOrH rH
^- rHOib-l>rí^ t-t-t> O O rH Ci 05
iO O O 00 lO
»O C l> lO CD
O O lO
g (Xi CO
»OOOOO
b- O ^ 05 TÎ<
CD (Tî CO CO (M 05 "^05 (M^ CO rH 05^ CO
'^ Ö f^ 4J "•^'
•"• a> 0) o ^ c6 oi (M" c<r
ap
r-> rO iO p lO o lO »o o o OOOOO
c^ t- CO 00 ^ CO rH CO 00 t^ lO 00 CO
'^^ '^ '^ ^^ "^ CO o^co i> CO CO^ K5 t^
»O TîT TiT ïo Ttn" co'~(ríoíoí(M"
OOOOO
TtH 00 CO T+( (M
CO^ rH o 00 05
iO Tfl Tt^ »o Tfl CO'cîC^'^rH rH
íís o o OOOOO
c^ o i> o t> CO rH CO -" 05 <N Tti T«
^ TtH (M TH CM CO O CO ^ rH CO t> 00
rH CM 05 05 CO 05 00 TH TfH CO
CM iO CO 05 rH O TH CM Tt^ CO
Ö ' • 'rH
fcJD Co
a^ n3 o
KÎ Q,
Pi ^ o P<
m 1^ tí
SOIL DYNAMICS IN TILLAGE AND TRACTION 339
the power requirements. These computations were made on the data
in table 42 for three hypothetical cases in which desired clod sizes of
14, 1, and 2 inches were selected. The evaluation plotted as a func-
tion of the tool capacity factor is shown for the three implements in
figure 236.
DESIRED MEAN
WEIGHT DIAMETER
.5 IN.
1.0 IN.
2.0 IN.
o
2
z
o
I-
<
z> /V
(A) (C)
CAPACITY (ACRES/HOUR)
7.1 Introduction
Traction may be defined as the force derived from the interaction
between a device and a medium that can be used to facilitate a de-
sired motion over the medium. The usual traction device converts
rotary motion derived from an engine into useful linear motion.
Anchor devices such as winch sprags are exceptions to the usual
concept, but they are traction devices since they provide traction by
interacting with a medium—usually soil. Although the basic soil
reactions are similar, the continuous rolling action of a wheel or
track requires a different analysis than the stationary action of a
sprag.
The traction and transport devices considered here operate off the
road and are construed to be wheels and tracks—that is, parts of
vehicles rather than complete vehicles such as tractors. The number
of off-the-road vehicles is rapidly increasing for agriculture, military,
and construction purposes. The total engine power available for
conversion into useful pull is generally in excess of the traction ca-
pacity that can be developed between the traction device and the soil.
In other words, the limitations of the vehicle in respect to off-the-
road movement are usually the limitations of the traction device.
Furthermore, the efficiency with which a traction device converts
energy into pull is usually extremely poor when the device is oper-
ating on soil. Work at the National Tillage Machinery Laboratory
shows that a pneumatic tire operating on a concrete surface has an
average power efficiency of approximately 75 percent. The same tire
operating on various soils has an average efficiency of less than 50
percent. Nebraska Tractor Test results indicate that pneumatic-
tired tractors operating on concrete lose approximately 5 percent
more in thermal efficiency when the useful work output is expended
through the drawbar than when it is expended through the power
takeoff. Obviously, loss in thermal efficiency on soil will be even
greater. Based on this minimum loss in thermal efficiency, some 152
million gallons of gasoline valued at $42 million are lost anually be-
cause of the inefficiency of the pneumatic tire.
There are soil conditions where adequate traction and satisfactory
efficiency can be obtained ; but, because of economic, social, and politi-
cal pressures, such tasks as pest control, crop harvests, and military,
mining, and construction operations are performed on extremely
poor soil conditions where adequate traction cannot be attained.
With the exception of loose dry soils, volcanic ash, or sand, most of
the adverse conditions are associated with wet soils. Some of the
land conditions on which operations must be conducted are so ex-
treme that entirely new principles of vehicle design may be required.
340
SOIL DYNAMICS IN TILLAGE AND TRACTION 34:1
^ä b
FIGURE237.—The traction force íT in a simple rigid body system as related to
the normal force V when the traction force originates from friction.
resulting from drawbar j)ull and from the torque reaction of the trac-
tion vehicle itself may significantly change the normal load at any
specific point in the contact area. Thus, alterations in the normal
load must be expected. Eegardless of the magnitude of the co-
efficient of friction, the limiting factor in traction where equation
153 applies is friction. The solution is very simple since the con-
cept implies rigid body behavior.
If the surface of the traction device is indented by sharp grains
of soil or if adhesion occurs between the soil and the device, inter-
locking may result. Under these circumstances, failure occurs not
by sliding friction along the soil-device interface but rather by shear
within the ground medium as shown in figure 238 (34,), When
failure is due solely to friction, failure occurs along the line /-&. For
a load 7, the force available for traction H can be increased until the
angle of sliding friction is exceeded. The force that causes failure
(sliding movement) is the maximum force that can be developed for
traction for that situation. With the equipment shown in figure 239,
the traction force that can be developed by simple traction devices
can be determined for varying H/V ratios. In addition, Bekker
observed through the glass side of the apparatus box an apparently
rigid soil body that formed and adhered on the bottom of the device
{fach in fig. 238). Under these conditions, failure occurred in the
soil rather than at the soil-device interface so that traction resistance
originated from cohesive as well as frictional components of soil
strength. The shear pattern that developed in the soil at failure was
considered analogous to the pattern that developed under an imagi-
344 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
///////////////////////
FIGURE 239.—A simultaneous loading device with which desired H/Y ratios
can be controlled in the study of traction by simple devices. (Bekker, Land
Locomotion Laboratory ( ^^ ). )
-CRACKS APPEAR
SURFACE MOVES
SAND
-CRACKS APPEAR
DISPLACEMENT
Y'i
Vo '
where v = velocity of the vehicle system,
Vo = initial velocity,
S = slip,
and by rearranging terms,
V = Vo (l-S), (155)
Equation 155 gives the velocity of the vehicle system—that is, the
vehicle velocity—with respect to the earth system in terms of an
initial velocity and slip. Using a unit vector i in the positive direc-
tion along the x and x' axes, the velocity of point A in the vehicle
system is
dx — — Svo dL
Jxo Jo
or
^x = x-Xo = - Svot, (159)
where t^x — the actual displacement of point A in the earth
system during time t..
In the vehicle system, point A is at some point x\ at time t = 0
and point A is at some other point x' at time t - t. If Z is a
measure of the distance between x'o and x\ then, since point A has
velocity Vo^ the time required for point A to move distance L is
V = 2bK \ i-^] dx =
Jo \L ) ""^ n+1
and solving for s„ gives
[ yiiLi^r" (162)
With So known from equation 162 and using the Bernstein equation
along with the relation between z¡c and Zo we can write
= K{^): (163)
which provides a relation between the distance along the track and
vertical pressure in terms of the vehicle parameters F, 6, and L and
dynamic soil parameters K and n. Figure 243, A shows such a pres-
sure distribution where n is less than 1. The average vertical pres-
sure on each quarter section of the track can be used with the Cou-
lomb relation shown in figure 243, B to determine the maximum stress
that could be attained by the midpoint of each track section.
Since not all points along each track section have been displaced
enough so that the maximum stress is attained, stress H versus dis-
placement j curves must be used to determine the average stress devel-
oped for each section. These relations can be conveniently measured
for loadings that correspond to the average normal load under each
track section (fig. 243 A) or be interpolated from a family of curves
350 AGRICULTURE HANDBOOK 316, U.S. DEPT. OP AGRICULTURE
P (p»i) ■ ■ ■
(A) lr>^1 I I I
(D) J <in)
that bracket the average normal loads. Figure 243, 0 shows such
¿^-displacement curves. These curves, and others for natural soils
in the field, generally appear to have the shape of type B in figure 41.
The curves representing average iï-displacement relations for each
quarter section of the track are shown separately in figure 243, D.
When the length of the track is known, equation 160 can be used
to determine displacement at any point along the track. Thus, the
portion of each ¿^-displacement curve between displacements at the
beginning and end of the respective track sections can be determined
for a given slip. Figure 243, D shows the appropriate sections of
the ¿i'-displacement curves that have been computed for the four
quarter sections of the track for slips of 10 and 15 percent. The
shaded areas beneath the appropriate portions of the curves may be
divided by the displacement range for each area to give the average
stress developed by each section. Multiplying the average stress by
the area of the track section over which it acts and then summing the
individual track sections gives the total traction force for each magni-
tude of slip. By using a number of slip values, a complete slip-trac-
SOIL DYNAMICS IK TILLAGE AND TRACTION 351
tion curve can be constructed for a given vehicle in a given soil con-
dition.
The foregoing procedures accomplish what must ultimately be
done mathematically. The graphical analysis approximates the tan-
gential stress distribution beneath the traction device by discontinu-
ous distributions. To be correct, the distribution must be represented
continuously so that an analytical correct summation by mathematical
integration can be accomplished. Janosi ( 198 ) has extended the
principle to a track and a wheel so that a rigorous solution can be
obtained. The method may give erroneous results, however, if the
equations representing soil behavior are incorrect. Thus, both the
Bernstein and Coulomb equations may give erroneous results even
though the mathematics is logical and rigorous. The method does,
however, demonstrate the essential requirements of a traction me-
chanics for a rolling device.
Until this point only methods of analysis that provide estimates
of traction force available to facilitate motion have been considered.
Not all of the traction force in a rolling device is available for useful
work. Some energy is required to deform the soil—that is, to over-
come the rolling resistance of the device. Thus, the net drawbar
effort will be the difference between the traction force and the rolling
resistance. Several attempts have been made to calculate rolling
resistance, but no method seems satisfactory at present (1965).
Eolling resistance has many visible forms : sinkage or compaction,
drag on the sides of the device, and a buildup of soil in front of the
device above the original soil level. Building of soil is often termed
bulldozing. These complex forms of rolling resistance are difficult to
represent by mathematical models, which partly explains the in-
ability to calculate resistance. When further considering that the
forces involved in rolling resistance occur in the same area as the
traction forces, it must be recognized that the traction devices do not
distinguish between traction forces and rolling resistance forces;
rather, the devices sense one distribution of forces. The concept of
two separate force systems is thus useful for understanding and for-
mulating the problem, but probably does not represent the physical
situation.
Steinbruegge ( Ji.ll ) proposed a concept of ideal traction efficiency
of soil that might provide a useful addition to a traction mechanics.
The concept employs the basic stress-deformation data, which can be
obtained from measurements of S'-displacement curves such as those
shown in figure 240. Since no traction force H is available until
some displacement occurs, no traction force is available until some
work has been done on the soil. The area under the ¿f-displacement
curve represents work that has been done on the soil since a force H
was moved through a distance j—the displacement.
Figure 24e3, D shows the average work that would be done by each
quarter section of a track operating at a given slip in the assumed
soil condition. The area under an ^^'-displacement curve is not the
only work done on the soil since, as sinkage occurs, the product of
vertical force and vertical displacement also represents work. Stein-
bruegge's concept is to maximize the ratio of traction force developed
to the total energy lost—that is, the work done on the soil. This
352 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
100%
IDEAL TRACTIVE
EFFICIENCY
0%—
TOTAL ENERGY LOSS
FIGURE 244.—The ideal traction eflaciency of the soil as determined from the
energy lost in deformation. ( Steinbruegge, 1st Internatl. Conf. Meeh. Soil-
Vehicle Systems ( ^ii ).)
mum. Steinbruegge concluded that the ideal efficiency gave the most
effective loading of the soil to obtain traction.
Obviously, an infinite number of possible loadings exists. A roll-
ing device does not first apply the vertical load and then begin a
horizontal displacement ; rather, the vertical load and horizontal dis-
placement are simultaneously changed. Therefore, families of H-
energy loss curves should be determined either at various constant
vertical loads'or at various rates of increasing vertical loads. The
maximum ratios attained by these families would represent the most
suitable combinations of loadings and displacements for the given
soil condition. From such information, a traction device designer
could determine the combinations of normal loads and displacements
that give the maximum ratio. As nearly as possible he could then
design the length and size of contact area and the distribution of
vertical load within the contact area to give the desired combinations.
In other words, the device could be designed from the standpoint of
the soil.
The ideal efficiency concept can be generalized even further than
indicated by Steinbruegge. He implied that the lost energy was
absorbed by the soil; hence, the term ideal efficiency of soil. In
reality, no assumption concerning the energy loss is necessary for
the concept to be accurate. The unit area may be visualized as a
rigid body; traction is obtained by the device thrusting against the
area. That the unit area is physically a part of the traction device
is not important. What is important is that the unit area moves
both vertically and horizontally, that this movement reñects energy
or work done on the unit area^ and that this work was necessary to
obtain traction.
The conservation of energy theorem states that the energy put into
the unit area must be either stored or transferred. Steinbruegge
considers the case where the energy goes into the soil. On ice, how-
ever, the energy can go into heat in the friction surface instead of
permanently deforming the soil and the approach is still valid.
Thus, in a generalized approach, the concept considers only the
movement of the unit area necessary to obtain traction. Where the
energy goes need not be considered. The device and medium are
considered as two separate rigid bodies and their combined behavior
as a system is described. The description might result in three-
dimensional plots rather than two-dimensional plots, as shown in
figure 244. For example, if H^ energy loss, and displacement were
shown in three-dimensional space, a constant vertical load would
describe some surface. The surface would represent the system be-
havior of the unit area represented (such as rubber, steel, or steel with
grousers) and the medium. For a different vertical load, a different
surface would probably result. Ultimately there must exist some
envelope of all possible surfaces, and this envelope could be maxi-
mized for various effects. For example, maximum 5^-energy loss
ratios could be determined for best efficiency of the medium or
maximum ¿"-displacement ratios could be determined for minimum
disturbance. Sinkage could replace displacement in the three-dimen-
sional space to obtain ratios where sinkage is important. To maxi-
mize S'-energy loss, the circular scale could be retained as a cylindri-
354 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
cal surface or perhaps other statistical techniques could be used.
Thus, by generalizing the ideal efficiency concept the behavior of a
system composed of a traction device and medium could be optimized
for a particular effect (maximum efficiency, minimum sinkage). The
conditions reflected at the optimum would represent the best condi-
tions at which to operate. Where some limitation restricts the range
of possible vertical loads, the optimum within the restriction can be
obtained. Where no limitation exists, the ultimate optimum (the en-
velope surface) can be obtained. The concept thus provides a method
for obtaining information useful in design. While much effort is
still required and many techniques still need to be developed before
practical results can be obtained, the potential of the generalized
ideal efficiency concept warrants the urgent efforts of future re-
searchers.
The assumption that rigid body behavior applies to both the soil
and the traction device provides a basis for developing a traction
mechanics. The assumption permits a development where stress-
strain equations of the soil are not required since the forces and
motions between the two bodies describe the desired behavior. While
an assumption of rigid body behavior eliminates any possibility of
describing behavior within the soil medium, it does adequately de-
scribe traction behavior. Thus, where only traction is to be con-
sidered, assumption of rigid body behavior is logical. Under such
an assumption, a description of the forces acting between the rigid
traction device and the rigid soil is adequate.
For a given device in a given soil condition, the method of Gamal-
ski ( Si ) demonstrated the principles of calculating the traction
force. Steinbruegge's concept goes even farther; he proposed that
an optimum loading can be determined for a given soil condition. A
traction device can then be designed to provide as nearly as possible
the desired loading for any situation. Ultimately, the methods could
be extended to optimize the design for a group of different soil con-
ditions.
Attainment of these goals is not at hand, but is a challenge to the
researcher. Accurate mathematical representations of the distri-
buted forces between a traction device and a medium are required.
A means of measuring ¿^-displacement curves that accurately repre-
sent the desired behavior or a means of calculating ZT-displacement
curves from dynamic soil properties is required. Better methods for
calculating rolling resistance are also required More accurate
methods for determining energy loss in developing traction forces
are needed. The attainment of these goals will provide a basis for a
suitable traction mechanics.
7.2.3 Transport Devices
A number of devices used to transport payloads over soil are not
powered; they must be pulled. These devices range from rigid
sliders such as sled runners to free-rolling wheels. The mechanics
of these transport devices is not discussed here. Skids or runners
operate by sliding, and the mechanics of sliding surfaces was dis-
cussed in section 4.3.1 The free-rolling wheel is a special case of the
wheel in which the torque is zero. The forces on towed wheels are
discussed later in this chapter (sec. 7.4.1.1).
SOIL. DYNAMICS IN TILLAGE AND TRACTION 355
TRACK PLATE
VERTICAL
FORCE
HORIZONTAL
FORCE
the soil. This dynamometer has been used to determine the forces at
all points of contact between a track of a vehicle and the soil (fig.
247). The force pattern applied to the soil may be determined from
such information.
One complicating factor is the direction of the applied forces.
When the ground is hard, the track is horizontal and the orienta-
tion of the dynamometer is known. When sinkage or bending oc-
curs, the orientation of the dynamometer in jiot known. Neverthe-
less, the lack of similarity between measured forces and those en-
visioned in figure 243 indicates the pressing need for additional
measurements.
AVERAGE
VERTICAL
PRESSURE
^FRONT
12 6 REAR
LENGTH (in)
FIGURE 250.—Pattern of pressure measured under the center of a smooth tire
operating in several soil conditions. (Vanden Berg and Gill, Amer. Soc.
Agr. Engin. Trans. {460 ).)
I I I I
8 4 0 4 8
LENGTH (IN)
FIGURE 251.—Pressure distributions under a smooth 11-38 rubber tire on firm
sand when inflated: A, To 14 p.s.i. ; B, to 10 p.s.i. ; C, to 6 p.s.i. The direc-
tion of travel was toward the right. (Vanden Berg and GiU, Amer. Soc.
Agr. Engin. Trans. ( 460 ).)
values at any location along the tire during one revolution. The
patterns, therefore, do not indicate the instantaneous stress distribu-
tions that were presented in figure 251.
The stress distributions along the surface of a traction device
thus are complex and may vary continuously as soil and operating^
conditions vary. Hopefully, a better insight into the significance of
these distributions will be obtained when improved instrumentation
IS developed. Knowledge of the orientation, location, and movement
of individual points along a tire will assist in providing parameters
that can be of value in describing tires.
The orientation of the stress transducers is usually not known
when flexible tires are used. Better instrumentation, however, can
help to overcome this difficulty. The same problem exists when
transducers are placed in the soil and large soil deformations take
place, since the transducers may rotate so that their orientation is
no longer known. Corrections, however, have been applied to these
position shifts {489), Transducers capable of measuring the tan-
gential components of stress as well as the normal components must
also be developed. Transducers will have to be constructed that will
react more nearly like the surface of the material in which they are
embedded. If the transducers are hard and rigid, they may cause
stress concentrations; or, conversely, if they are soft, they may cause
a bridging over the transducers. In either case, inaccurate measure-
ments will result.
In the past, because of the fragile nature of the sensing devices,
measurements have been made only on slowly rolling, lightly loaded
wheels. The strength and durability of these devices will have to be
increased so that measurements can be made on flexible wheels hav-
ing high torque inputs and high vertical loads. Both types of
loading cause distortions and deflections in tires. Experimental evi-
dence indicates that the distributions will differ not only for different
soil conditions but also for different tires and for different degrees of
wear on a given tire. In each case the distributions will have to be
measured, after which generalized descriptions of stress distributions
between the traction device and the soil may be made. The extent
to which these descriptions can be placed in mathematical form will
determine their value in contributing to a rigorous calculation of the
traction capability of traction devices.,
7.3.2 Deflections or Movements Between Devices and the Soil
The distortion of pneumatic tires is one of the complicating factors
in the study of their action on soil. Since flexing permits the tire
to act as a nonrigid body, the direction in which stresses are applied
and the size and shape of the area of contact may not be known.
As a tire rolls over a surface, squirming or rubbing along the sur-
face of the tire results from deformations within the tire. The im-
portance of these movements on traction and wear of the tire has not
been fully examined because the movements have not been completely
measured.
Several techniques have been developed for exploring these move-
ments. Cegnar and Fausti ( 65 ) measured the movement of fixed
points on a tire with optical equipment as the tire rolled over a flat
surface. The actual movement or slippage of the lug at the center
SOIL DYNAMICS IN TILLAGE AND TRACTION
movement of bars 1 inch wide when a tire was rolled over the bars.
Notice that the central section of the tire caused the bars to move
forward while the exterior sections caused the bars to move rearward.
Whether such measurements can be used to characterize the traction
performance of tires has yet to be determined.
External deflections recorded by scratch plates or other techniques
364 AGRICULTURE HANDBOOK 816, U.S. DEPT. OF AGRICULTURE
_ '40» 40»
0« 0« 0«
//////
/////
propelled state. Data in figure 257 show that the rolling radius
(the distance the axle moved forward divided by 2 TT) for the condi-
tion (2,0704b. load at 14 p.s.i. on concrete) was not uniquely related
to either the axle height or the undeflected radius (no load on the
tire). If the tires are operated in a soil where smkage occurs, a
different rolling radius probably will be found at the self-propelled
point. Quite clearly, no simple method exists for defining a theo-
retical rolling radius of pneumatic tires. . -...r
The deformation of flexible traction devices takes place m diöer-
ent directions {m), as illustrated in studies by Wann and Eeed
( IL70 ) The relative movement between the contact surface o± a
tire and a hard surface was studied by a flat scratch plate technique.
Carborundum particles were uniformly distributed over a polished
aluminum plate after which a tire was loaded and towed over the
surface. The carborundum particles scratched the surface ot the
plate as they were moved by the scrubbing or squirming of the tire.
A special red-base waxed paper was bonded to the surface ot the
plate to record the scratches. Figure 258 shows the scratches pro-
duced by a tractor tire having two loads and inflation pressures.
The paths along which the individual grains of carborundum moved
show clearly that the direction is sidewards as well as rearwards in
some mstances. In addition, the scratches reveal one part of the con-
tact area may be operating at essentially zero slip while another part
has slipped considerably.
The following explanation for the action was given by Wann and
Eeed {^70). The circular cross section of the tire can be con-
sidered to be made up of an infinite number of circles of varying
circumference, which are constrained so as to rotate on one axle
and operate as a unit. As a result, differential slip between the
various circumferences and the surface must occur. To examine
this action of tires, a bar table consisting of flat bars 1 inch wide
operating on and between frictionless bearings was used. The bars
were so supported that they were free to move lengthwise even
though they were kept in alinement and restrained from lateral
motions. The use of the bar table permitted the measurement of
the cumulative relative movement in increments 1 inch wide as a
tire traveled forward on the bars in a self-propelled state (fig. 254).
Figure 259 shows the result of the measurements for several ex-
perimental tires. The data were expressed in terms of movement
relative to the center line of the tire per unit distance of travel and
show the same general effect indicated by the scratch tests.
The complexity of defining zero slip for a flexible traction device
becomes apparent in the light of the numerous movements indicated
by the data just discussed. The problem is to define, the theoretical
distance or velocity traveled at zero slip. Since no convenient method
based on relative movement is available for describing zero slip, an
alternate method is to describe the state of forces on the device. For
example, in the theoretical wheel operating at zero slip, the torque
and the pull are simultaneously zero. In the real wheel where
rolling friction (rolling resistance) is present, torque and pull are
never simultaneously zero. Phillips ( 335 ) suggested that three
methods of operation exist on which a zero slip measurement might
be based: (1) the distance traveled by the wheel in one revolution
at zero torque; (2) the distance traveled in one revolution of the
wheel in the self-propelled state; and (3) the distance traveled when
the instantaneous center of rotation of the wheel is at the undis-
turbed surface of the medium on which the wheel is operating. All
three methods of operation would predict the same rolling radius
for a theoretical wheel if no sinkage occurred.
When sinkage does occur, the third method does not seem to be
logical since it results in predicting a rolling radius that is less than
the actual radius of the wheel. Figure 260, A shows a deformable
wheel operating in a medium where sinkage occurs and B^ on a
medium where all of the deformation takes place within the wheel.
Within the contact area, the radius of the wheel varies from R^ to
Ro; however, the value R^ depends on the soil condition and the
character of the wheel. The rolling radius should logically lie be-
tween these tw^o extremes. The data in figure 257, obtained with a
self-propelled wheel having forces as depicted in figure 256, />,
SOIL DYNAMICS IN TILLAGE AND TRACTION 373
liJ
.03r
X
o
er
Lu
CL
UJ
-.01 12 3 4
(^ DISTANCE FROM CENTER LINE (in)
FiGUBE 259.-Accumulated movement of bars relative to the center of the tire
caused by differential movement of different sections of the tire: A, conven-
tfonalDlv smooth tire; B, conventional-ply lugged tire; G, radial-ply smooth
tiref I! rfSatply lugged tire. (Reed, National Tillage Machinery Labora-
tory.)
'm^//À^/y
(A) (B)
TT^;;' (165)
where Vt is the theoretical velocity, v is the actual velocity, and P
and TF are the pull and weight on the device as indicated in figure
245. Both V and Vt are a function of rotational velocity co; and in
equation 155, vt is the same as Vo for a constant w. Equation 165
can be made independent of rotational speed by substituting for
V from equation 155, giving
^(1-^). (166)
where 8 is slip.
In equation 165, since the rotational velocity co can be maintained
constant, the denominator (Wvt) will be constant. The numerator,
however, >aries from zero at zero slip (since P is zero through a
range of positive values) to zero at 100-percent slip (since v is zero).
The units of the numerator are those of work so that the expression
represents the work output of the device for certain fixed conditions
(weight and rotational velocity). The expression thus may be called
a work output, and it is an index that represents the work in di-
mensionless terms. Since the work output reaches a maximum, the
pull associated with the maximum represents the pull at which
maximum work can be performed. The denominator in equation
165 IS not work even though the term has the units of work. The
vectors Vt and Tf operate at right angles to each other so that their
product is not work in the normal sense even though a superficial
inspection would so indicate. Because the work number in equa-
tions 165 and 166 represents work done, the number is a useful
criterion of traction performance.
7.4.7.4 ioocf-carr/ing Capacity
Transport performance may be evaluated in terms other than
towing force and speed. For a self-propelled vehicle, the time
required to move a specified distance could be a criterion of per-
formance. With rare exceptions, however, the object of vehicular
travel is to move a load rather than merely to deliver the operator
from one place to another. Dickson ( 102 ) proposed a load-carry-
ing index which was suited to a vehicle with a drawbar load ; a more
general form might be
SOIL DYNAMICS IN TILLAGE AND TRACTION 377
where Wp = payload,
T — torque,
CO = angular velocit}^,
P = drawbar pull,
V = velocity.
For a self-propelled wheel, the pull term Pv would go to zero. The
advantage of this form of load-carrying index is that it also applies
to the transport wheel. In the latter, the torque term Tœ goes to
zero and P is a minus quantity signifying a towing force. Wp
would be the load transported on level soil. If the vehicle climbs
a hill, the force parallel to the ground line will increase. The in-
crease acts as additional drawbar pull P which causes a decrease m
the load-carrying index when the vehicle operates on slopes ( ^/5 ).
Load-carrying index is thus a useful criterion for evaluating trans-
port.
7.4.2 Measures of Performance
As indicated in section 7.4.1, the first step in evaluating traction
performance is to establish fundamental criteria of performance.
Since traction devices must be compared, the second step is to quan-
titatively measure the established criteria. The criteria can then
be compared. For example, in one instance the most important
criterion may be maximum pull, whereas in another instance it may
be power efficiency. These fundamental criteria are not independent,
as has already been indicated (sec. 7.4.1.1), but are interdependent
in various complex relations. Furthermore, the relations depend
on the state of operation of the device. Therefore, to establish these
relations, the fundamental criteria must be simultaneously measured
in various states of dynamic equilibrium. Eegardless of the funda-
mental criteria chosen for comparison, certain basic measurements
are required. Pull, torque, weight carried, rate of angular rotation,
forward speed, and slip are basic measurements that describe the
state of dynamic equilibrium. Composite criteria such as power
eificiency can be calculated from the fundamental criteria, bpecial
equipment is generally required to control the operating conditions
and to make the measurements. The equipment shown m figures
2 and 3 was designed for this purpose. Figure 261 shows traction
measurement equipment at the National Tillage Machinery Labora-
tory. The control and measurement of fundamental criteria with
this type of equipment provides reliable information that can be
used to evaluate traction performance.
Traction depends on the relative displacement ot the traction
device, as demonstrated by the ¿T-displacement curves shown m
figure 240. Equation 160 shows the relation between slip and rela-
tive displacement for a rolling device; this relation was discussed
in detail in section 7.4.1.2. Figure 262 shows typical traction data
where torque and pull have been plotted against slip by using a
self-propelled state as the zero slip condition. Data on traction
378 AGRICtLTURE HANDBOOK 316, U.S. DEPT. t)F AGRICULTURE
:SS¡BS^S^¿
- BRAKED DRIVEN -
in figure 262 thus represents the operating states where the device
is providing a draying force and hence is a traction device. As
figure 262 shows, however, the device can be used for other purposes
such as braking and the total information shown is useful and
characteristic of both the device and the medium on which it is
operated. n - j-
Because of the mechanics of some types of equipment tor measur-
ing traction, weight may be transferred to the traction device as
drawbar pull is increased. Weight also varies with the slope of the
ground. These weight changes must be measured and the traction
results expressed in a form independent of the changes. Such a
form is provided by the coefficient of traction, which is defined as the
ratio of pull to the total weight carried.
Use of the coefficient, however, requires the assumption that the
coefficient is independent of the normal load. Quite clearly, such an
assumption is not justified when large variations m the vertical
load occur. For example, if a coefficient of traction-vertical load
relation is visualized, the pull capability of the device reaches a
finite maximum; and as the vertical load increases, the coefficient
must approach zero. At the other extreme, in the limit, the co-
efficient becomes undefined since it approaches the ratio zero/zero;
however, researchers ( lOß ) indicate that the coefficient approaches
infinity as the vertical load approaches zero. Approaching infinity
does not seem unreasonable; it merely indicates that the rate at
which pull and vertical load approach zero are different and the
rate of pull is slower. The important point is that the coefficient
does not express the desired information, namely, the maximum pull
for varying normal load; consequently, when trying to express the
effect on pull of varying the normal load, the coefficient of traction
is meaningless. When, however, small changes occur m the normal
load (such as caused by weight transferred withm a vehicle or
added from mounting an implement on a tractor), the coefficient
is necessary in spite of its dependence on normal load.
7.4.3 Evaluation of Performance
Evaluation of traction devices involves comparing the performance
curves of different devices with subjective levels of performance that
are required or desired. Devices may also be evaluated m respect
to each other without specification of any level or standard of per-
formance. Unfortunately, this type of comparison provides no com-
mon denominator for other evaluations. The difficulty m comparing
complete performance curves such as shown in figure 262 is that
they usually have different shapes for different traction conditions,
and the curves for two similar devices will often cross m the same
traction condition. Thus, for example, tire A may pull more than
tire B at 10-percent slip, but the reverse might be true at 50-percent
slip. The difficulty of comparing th^ curves can be greatly reduced
if a single value can be obtained to represent the overall nature ot a
particular curve H61). . ^^ ^^ ^ u
One value which might be used is the maximum pull that can be
developed. When a heavy load needs to be started or when a short
distance is to be traversed in difficult trafficable conditions, maximum
pull is perhaps the most important consideration. Maximum pull is
380 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
800 r
2000
LOAD (Lbs)
FIGURE 263.—Towing force required to roll various loads over a plowed loani
soil : A, On a 6-28 steel wheel ; 7?, on a 6-16 pneumatic tire ; C, effectiveness
of the pneumatic tire in reducing the towing force. (McKibben and
Davidson, Agr. Engin. {210).)
12 16 20 200
Y (lb/in)
where d diameter,
<f) (172)
b width,
F towing force (rolling resistance),
W weight on w^heel,
K proportionality factor reflecting soil conditions.
Based on Freitag's analysis, data from McKibben and Davidson
( 270 ) were analyzed by Gill and Vanden Berg as shown in figures
265 and 266. The relations indicate that the trends predicted by
equation 172 apply to conditions other than loose sand ; the reaction
to loads on concrete and bluegrass sod is considerably different than
on sand, yet the relation holds. The points representing two of the
six wheels reported in figure 266 (shown as circled crosses) appear
to be on a different line than the remaining four wheels. Since the
measurements of these same wheels did not appear to be abnormal
SOIL DYNAMICS IN TILLAGE AND TRACTION 389
3600
3000
2400
1800-
1200
600
X 1000
1200
1000-
800-
600
Ü.
400-
200
in the other test conditions, some change may have occurred in the
sod while the measurements were being made. The circled points
may thus reflect a changed sod condition rather than scatter in the
measurements.
Figures 265 and 266 also show that the proportions between the
factors m equation 172 may not be as simple as first indicated,
bmce a best-fitting line does not pass through the origin, equation
172 has the form
17 \ 3/2
^=.( T) +^' (173)
where C = SL constant.
Equation 171 suggests that two dynamic properties, which are com-
mon to the steel wheel and the medium, describe the behavior. Ke-
gardless of the implications of K and G, the effects of composite
design factors can be demonstrated by rearranging terms in equation
173 to give
where the terms are the same as those previously defined. Thus,
increasing the diameter decreases the towing force and increasing
the load on the wheel increases the towing force. But the effect
of width IS not clear. Increasing the width decreases the left arm
m the bracket of equation 172 but increases the right term. For
values of i at infinity and zero, the quantity in the brackets becomes
mñnite so that some minimum value must exist at intermediate
values. Taking the partial derivative of the quantity in the bracket
with respect to b and equating the derivative to zero provides a
method for finding the minima and gives
1 5rTF3/2
2 53/2 ^ ^ - tF,
^ = (^) W, (175)
which indicates the value of b that gives the lowest towing force.
The effect of the shape of the wheel on the towing force cannot
be determined from available information. The wheels considered
by Freitag and by McKibben and Davidson had flat rims. One
isolated comparison of different-shaped wheels indicated that both
convex- and concave-shaped wheels required a slightly larger towing
force, on the average, than a flat wheel of the same width and di-
ameter. While the convex wheel required a larger towing force
than the concave, the difference was small. The effect of shape
appears to be much less significant than the effect of either diameter
or width ( ^7^, 274 ). If a more exact evaluation of the effect of
shape IS required, additional research is needed.
Equation 174 can be shown to be generally applicable to pneumatic
SOIL DYNAMICS IN TILLAGE AND TRACTION 391
tires. Freitag demonstrated that equation 172 applies if tires op-
erating on loose sands are compared at constant deflection. Thus
for a varying load, inflation pressure or size of tire must be changed
in order to maintain a constant deflection. Deflection thus appears
to be a measure of the flexibility of a pneumatic tire. Figure 267
wheeled vehicle. The towing force required for the latter was at
least 100 percent larger in nearly all instances. The performance
results might be reversed if the two vehicles were operated on roads
at higher speeds. Eecent developments in rubberized flexible tracks
may overcome this limitation.
PULL (Lb)
50
5 X 12 DIAGONAL
12 LUGS TOTAL
4 X 22 DIAGONAL
g 40
Ö
5 30
4X9
4X6 SWkDE
SPADE I
20 -> ^
16 LUGS TOTAL
10
PULL (Lbs)
\ BLUEGRASS
SOD
UJ
o
\i¡ \ PULVERATED SOIL
ÜJ
il 13 15 II 14 17
PULL X 100 (Lb)
9 12
PULL X 100 (Lb)
decreased, spacing between the lugs increased. The data indicate that
power efficiency increased as lug spacing increased. Until a condi-
tion is encountered where grouser action is necessary to obtain in-
creased pull, presumably the trend can be extrapolated to the condi-
tion where no lugs are present and the tire becomes a smooth tire.
Technical difficulties during the conduct of the measurements were
believed to have contributed to some of the erratic results m figure
272. The dotted extrapolation of the curve for the 26-lug tire and
the apparently different behavior of the 23-lug tire m the clay soil
cannot be fully explained. ^ . • i. ^
The data reported here do not indicate what effect lug spacing had
on maximum pull. Since power efficiency approaches zero as slip
increases, the respective curves shown in figure 272 do not give a
reliable indication of the maximum pull that was developed. Fewer
lugs, however, appear to contribute to increased power efficiency. It
is possible that reducing the number of lugs increases power efficiency
for the same reason postulated for reducing lug height. Since the
soil is disturbed less with fewer lugs, less work has to be done on the
soil to obtain traction. This line of reasoning suggests that when
tires develop comparable pull, the tire that disturbs the soil least
should be the most efficient. If such a conclusion is valid, the obvious
interaction of lug height and spacing becomes apparent and indi-
cates that they are composite design factors.
398 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
3iOr
Maximum Average
Design coefficient of coefficient of Average power
tractioni traction^ efficiency^
factor
Lugged Smooth Lugged Smooth Lugged Smooth
Percent Percent Percent Percent Percent Percent
Smooth (lugs
removed) -3 26.0 24.0
Radial ply
construction-. 0 1 15 10.0 4.0 2.0
Narrow rim 2 2 6 4.5 4.5 -0.5
Flattened
tread base -1 0.5 1.5 0.5
Total improve-
ment of com-
bined factors. -1 26 41.0 10.0 26.0
1 Determined in the range of 0- to 70-percent slip.
2 Determined in the range of 0- to 30-percent slip.
3 For coefficients of traction in the range of 0- to 30-percent slip
SOURCE : Vanden Berg and Reed ( 46i ).
cient in energy transfer and will even pull more at lower slips. This
confirms again the results of the lug-height study (fig. 271) and
lug-spacing study (fig. 272) for pneumatic tires and the lug-height
study (fig. 269) for rigid wheels.
The fundamental prmciple of driven wheels is clearly demonstrated
by the various studies of lugs. Except for maximum pull, no lugs
are needed and their presence reduces power efficiency. The de-
signer, thus, is faced with a compromise. Where poor traction con-
ditions are anticipated for the major use, an aggressive lug may be re-
quired and justified at the expense of power efficiency. Where rea-
sonably good traction conditions are anticipated, only a minimum
amount of lug should be used so that power efficiency is kept high, yet
a reasonable maximum pull can be attained. The information re-
quired by the future designer is the magnitude of grouser action
needed to obtain a desired increase in maximum pull for the ap-
propriate traction conditions. With such information he can effect
the appropriate compromise.
A second conclusion concerns the effect of radial-ply construction.
The design factor had no significant effect on maximum pull in either
lugged or smooth tires. However, a significant increase was obtained
in average pull and power efficiency. Since the increase was re-
flected in both smooth and lugged tires, radial-ply construction
must affect the manner in which forces are transmitted from the
rim to the soil so that traction performance is increased. The
factor, thus, is a carcass factor and suggests that considerable
knowledge is needed in two areas: first, the distribution of stresses
required by the soil for optimum traction; second, the manner in
which carcass construction affects the transmission of forces from
the rim to the soil-tire contact surface. The present state of knowl-
edge indicates that radial-ply construction will not effect maximum
404 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(A) (B)
FIGURE 277.—Area of soU disturbed: A, By a single large lug; B, by several
small lugs.
406 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(A)
(B)
(C)
COEFFICIENT
I 2 3 4 5 6
of the vehicle reduced the maximum efficiency and the pull at which
it occurred. Eeed ( 35Jf ) studied the influence of varying the load
by means of changing the width of the track. As shown in table 49
increasing the normal load on a clay soil increased pull. In sand,
little influence was detected.
FIGURE 282.—The forces on an improved winch sprag designed to load the soil
faUure surface: W = weight of tractor, A-B = area of slip, s = shearing
stress Ô = angle of soil-metal friction, T = maximum traction force, a -
angle of failure (7r/4-0/2), R = resultant on sprag, <T = normal stress
on surface, R, = resultant of soU. (Payne, Jour. Agr. Engm. Res. (SSO).)
SOURCE : Payne (.
410 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
pensions may be nonrigid, which permits a degree of vertical flexi-
bility ; but each track shoe remains inflexible. Forrest ( 123 ), speak-
ing for the American Eubber Industry, has indicated that rubber
tracks offer great promise for agriculture since they provide low
ground pressure intensities, are lightweight, and may be operated at
much higher speeds than vehicles with metal tracks. A rigid high-
speed metal track has also been developed.
Bonmartini ( ^9 ) has designed a number of pneumatic flexible
tracks for aircraft and other vehicles. Although the specifications
and techniques of construction of flexible tracks are beyond the
scope of consideration here, these are the factors that have slowed
the development and use of flexible tracks. One of the more radical
designs is a roller track. The individual track shoes are H-shaped
and contain a roller in each of the open sides. During operation,
the center bar of the H is perpendicular to the direction of travel
and provides traction as do the rollers whose axes of rotation are
mounted parallel to the direction of travel so that they do not roll.
On turning, however, the lateral movement of the track is facilitated
by the free-rolling movement of the rollers on each of the open
arms of each H-shaped track shoe. The roller track can turn with
only 50 percent of the torque required for rubber tracks and about
30 percent of that required for a steel track. Lateral movement on
slopes is reportedly not a limiting factor of the design, and the
maximum pull that can be developed on level land equals or exceeds
the pull that can be developed by steel and rubber tracks. Slope
stability has been increased by introducing an angle of 10° between
each of the rollers and the center line of the track.
Tracks will continue to be designed through the use of composite
parameters. These parameters appear to be identical for tracks and
wheels, when allowing for variations such as the relation of the
radius to axle height. Other aspects such as vibration and soil com-
paction have been studied, and additional design information con-
cerning these points is included in chapter 8.
rS 1 Ol Pi ^•o coa t-
r«<i
S:^.
a8 r—i\ÔZJ"~* o X r-<
r-i
9^ S'^ ÛH
■1 si a (1^ ft
0) g ^:;
d 4) îi? '-gg
^ ■^ ft 1aa
<2^
>^ 4J -M
1 8 8^^
GO 3^^ ä «> vA TH (N C^
^^<o (1^ ft
53
1- a «
S g g 8 -M
r-î rAoirA
r< ^ 00
-M -»-3
5^ ci Ö ^ O CD O
8 O t' ce
Co" 3^ S'^ TH r-î T-î ci
.2 a p^ ft
Î ^1
S-"
§^l
ce
gÍ3 8 î:5g8 O Cj O
S a» TA r-î r-î C^*
1 CÖ ^ ..
•§
't:^
-4J
3^ ë^
+J
8 8g8
r-î rHr-îc4
pi
rf o «2
^ PH ft
11 « o
^2s
ga^^
8
T-Î
p looq
T-î r-î r-î
+-> 4J
CO CD 00
8 O iO CO
a« Ci^ ft
rH rH (N
'-tó %n
s- 8 ë^S
;^cá 2a »
s T-Î r-î r-î TA
•^ -M +J
1 a; ft
8 05 ÍÍ5 00
3^
pLi
Sft * r-î * r-î T-î
«ï)
1 cj 2 í3 8 8g§8
gaa TH r-î r-î r-î
•iS>
<ï) 1
1 03 Cj 02
1
-S 1
i 1 i i o'^ a
2
*o3
5 ' 1 1 P-H§12
O ft 1 1 1 o3 bu
°f 1^
d
o Oí
^P
02 œ
?>
1
1
1
1 02
1 'd'ö
o a>
-M +-»
d
0)
(D
'-M
4^2 S i 1« 03 c3 Wi
13 *-M *4-> 4J
.è ,^Qa2W r-l (N 00
H 1 <i<^
412 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
FIGURE 283.—Wires used to cut mud from a rolling wheel to prevent im-
mobilization of the vehicle.
SOIL DYNAMICS IN TILLAGE AND TRACTION 413
less than the rubber tire to which they are attached, they will not
touch the highway.
Traction aids may be designed to improve the self-cleanmg quali-
ties of devices. For example, figure 283 shows the use of short
lengths of tightly stretched wires to cut mud from a rolling wheel.
While cutters or scrapers may prevent an undesirable buildup of soil,
they require an additional input of work to remove the soil. Use of
polytetrafluoroethylene or other materials on devices may prevent
soil buildup by reducing the adhesion of soil to the device.
7.5.5 Operational Control of Design Factors
A number of design factors can be modified, adjusted, or controlled
by the operator to improve the traction or transport performance of
a vehicle. Although the exact nature of their influence may not be
known, their immediate effect on performance can be observed and
adjusted. In a final analysis, the operator should be able to adjust
all parameters to meet individual operating conditions. The simplest
example of this type of control is altering the inflation pressure of a
pneumatic tire to provide different tire deflections and changes in
rolling resistance {271), Decreasing the inflation pressure may
improve traction performance in some operating conditions. As
shown in figure 284, the pull of a tire may be increased by deflating
^'205 4ÔÔ 600 8ÖÖ IOC» 500 1000 1500 2000 2500 3000
4J -M
^H lO o JPH ^
8 (M 22
o
CO
ÍO
rH
y^
3^ g'^ TH TH TA rA ci
P-i a
Ö2 ,
<.ñ ■Sis
OS i2 i3 8 Ci ^
^ a i=^ T-î TH TA vA oi
8 rH ^ <N 00
>j
i3^ g CO TH TA
piH a
a '->
gÍ3 8 S 8 g ^
gaa TH r-Î rH rH C^*
o .-s
8 CO 8 8 S 2a-
a, TA rH rH (N
.3 a
g" 8 Ci
!§ S?
. 4-Í
CO ?1 O 5 O
^ a a T-H íH TA TA (N
,CJ Q ..
4^ 4..
ci fl ^
8 8 CO rH
a« ill a
TA rA (M*
.2 a
■gg
8-^ 8 00
S ^ S?
gÍ3 CO
^aa TH r-í rH* rH rA
ci 4-2 4->
1
1 a> a
8 1
l> Ci iO
a« 1 TA rH C^
.sa
•tí ai
fil a
5 S
S-^ Mgö
3 s í3 S
1
1
ÏO
TH
rH
-^
t-
CO
ce ^ a ft TA
' TA T-î Oi
4^ 4^
8 ÏO
o
t-
Ci
rH
^
;0
CO
T-î TA rH* <N*
(>>iH
(^ a
1^
SÍ3 8 s s 5 3
lao. TH rH rH C4
TH
t3 . ^
1 1
1
1
1 pas
! 00 O Ö
c^
Oí 4-2
Á ó
73 a>
^1.^
00 00 cô
S fl 2
ga 25 (N <?;l S
M
2^ 3 rH rH 0
'S 'Ö Ö
0) Oí 0)
4J jj ûj
^ ±î '-0 4.Ï
CÖ rO Oro
çfl CO ce
ti ^^ ti
nu'î
-^ 3 CÖ- s|s
1 ^ < i-q 02 m
SOIL DYNAMICS IN TILLAGE AND TRACTION 415
the inflation pressure except for completely filled tires. The infla-
tion pressure in contact area of the tire may also be altered by
changing the size of the area of contact—that is, by using larger
tires or even by using dual tires. Adding a wheel having design
features identical to those of the basic traction or transport unit on
a vehicle is not considered a traction aid. Instead, it is considered
that the available designs are being exchanged or regrouped by the
operator. As shown in table 52, the traction performance of a ve-
hicle may be improved considerably by this technique. Dual tires
may increase the average pull of wheels more than 30 percent above
that secured by a single tire.
Since the source of the load on the traction device is not significant,
it may be obtained from a component of the pull as well as by adding
wheel weights. Implement hitches have been designed to transfer
a component of pull into a vertical load on the tire {JßO), A
vertical force is automatically^ added to the wheel as the pull is in-
creased, and additional traction is obtained. The hitch is a link
between the forces on a tillage tool and the traction devices re-
quired to pull the tool. Knowledge of the soil is obviously required
to relate these two sets of forces in any quantitative terms.
In sticky soil, the lugs become filled and lose their ability to cut
through the surface conditions that limit traction. Tires with di-
rectional treads may be fitted on the vehicle so that their move-
ment in the soil tends to scour the soil from the lugs. Figure 285
o a
I- -I
o (O
LÜ
£ Ü.
5 o
(B)
shows two ways that lugs may be oriented to the direction of travel.
In figure 285 J., wheel slip packs soil between the lugs and reduces
their effectiveness. But in sand or gravel or other conditions where
there is no adhesion, fitting tires as shown in figure 285, A may pro-
vide greater traction rather than less. Traction devices may also be
cleaned by reducing the inflation pressure to the point where flexing
is sufficient to assist in dislodging compacted soil.
These few examples show how the operator can control design
factors. The examples also indicate the need for a fundamental
mechanics that will enable intelligent control of the various factors.
416 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
Changing the height of hitch (fig. 280) for tracked vehicles is an-
other example. In four-wheel-drive tractors, a change in load dis-
tribution occurs with a change in height of hitch.
The operator will need simplified instructions or a simplified means
of evaluating a situation in order to make an intelligent decision.
Obviously, the operator cannot justify running through a series of
lengthy calculations to obtain information on which to base a de-
cision. All design factors that are subject to operational control
must, therefore, be translated into simple relations in order to obtain
the best traction possible. Very little information is available in
such a form today.
While no particular point has been made of operator skill, it is
obvious that the actual manipulation of vehicles is of paramount
practical importance in securmg traction and transport in difficult
soil conditions.
^sontact weight
pressure X factor
Mobility = 0.6 factor
+ wheelload
index tire X grouser
. factor factor
trans-
clearance X engine X mission + 20, (177)
factor factor factor
where
gross wt. (lb.)
contact pressure factor = tide width X rim diam. X No. of tires'
weight factor: greater than 35,000 lb. = 1.1
15,000 to 35,000 lb. = 1.0
less than 15,000 lb. = 0.9,
tire factor = 1.25 X tire width in inches divided by 1.00,
grouser factor: with chains = 1.05
without chains = 1.00,
wheel load = gross wt. in kips (wheels may be
No. of wheels single or dual),
clearance factor — clearance in inches,
10
engine factor: greater than 10 h.p./ton = 1.00
less than 10 h.p./ton = 1.05,
transmission factor: hydraulic = 1.00
mechanical = 1.05.
The mobility index is a specific characterization of a vehicle since
it includes only vehicle parameters. Indices have been developed for
self-propelled tracked and wheeled vehicles and for towed tracked
and wheeled vehicles. Empirical correlations have been made be-
tween the mobility index and vehicle cone index measurements.
Vehicle cone index measurements (sec. 3.3.1) were made on soil to
determine the minimum strength required for 50 passes of the vehicles
in the critical soil layer (6- to 12-inch depth zone). The mobility
indices of new and uncharacterized conventional vehicles have been
predicted with considerable accuracy by means of these relations so
they are of practical significance. Based on empirical relations such
as these, it is possible to analyze the morphologic factors that have a
pronounced influence on vehicle performance. This analysis may
be a means for detecting objectives for improved designs.
7.6.2 Vehicle Capabilities
The capabilities of vehicles may be varied by different designs to
meet certain forseeable requirements. A common example of this is
418 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
the construction of all-wheel-drive vehicles. Powering of all wheels
not only may eliminate the rolling resistance of any towed wheels but
also may add to the traction force that can be developed by the
vehicle ( 156, 358, 504),
A consideration of power efficiency-slip relations on different soil
conditions might indicate that the leading and following wheels on a
vehicle should be operated at different slips to achieve maximum
efficiency. A form of controlled wheel slip is achieved by wheel
braking or by using differential locks. Data reported by Geiger
( HI ) and Seuser ( 38^ ) indicate that use of a differential lock
prevented undesirable slip when plowing with one wheel in the plow
furrow. Control of the slip eliminated the need for wheel weights.
Individual motors on wheels provide another means of obtaining
controlled wheel slip. Depending on need, all-wheel drive may be
used to improve traction performance without adding weight or
using other traction assistance such as winches or tow vehicles.
These last two methods of securing propulsion recognize that the
traction capacity is inadequate and they also consider that the re-
quirement is very infrequent. Thus, use of a tractor to push a large
earth-moving machine while cutting and loading may represent a
more practical means of supplying power than to redesign a larger
and more powerful unit that will be used to capacity only on an in-
frequent basis.
Gross flexibility can be designed into vehicles in a similar fashion.
Articulated vehicles permit a measure of steering and climbing cap-
abilities that cannot be obtained by orthodox vehicle designs. The
Meili Flex-Trac is a six-wheel vehicle with a vertical articulation cap-
ability. Since any pair of wheels can be elevated, the vehicle can be
adjusted to conform to rough terrain so that the traction capacity of
all wheels can be brought into use. Horizontal articulation facili-
tates the steering of extremely long vehicles, especially when tracks
are used.
Kereselidze, Khukhuni, and Shkolnik ( 216 ) and Lange ( 250 )
have developed hillside tractors that can redistribute the load of the
vehicle on slopes to increase traction and stability. Soehne ( 400 )
has used a rolling disk to stabilize vehicles on slopes.
One novel capability that can be added to vehicles is lift. The in-
clusion of a fan to create a ground effects machine may reduce the
axle weight or completely lift the vehicle in nontrafficable conditions.
At the point where the complete weight of the vehicle is balanced,
the vehicle is no longer a ground-operating vehicle.
Special vehicle capabilities must be designed into the vehicle.
That is to say, information must be made available to designers to
provide a basis for the designs. The value of any particular design
must be based on a realistic assessment of its importance and fre-
quency of use. Some optimum balance between expense and need
would have to be reached to determine when any particular design
would be practical. These designs are then developed to exploit the
traction capabilities of the individual traction and transport com-
ponents of the vehicle.
SOIL DYNAMICS IN TILLAGE AND TRACTION 419
10 20 30 40 50 60 70
SLIP (%)
FIGURE 286.—Relative effect of traction conditions and traction d^ice design
on the performance of two 11-28 pneumatic tires. (National TiUage Ma-
chinery Laboratory.)
420 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
the rather radical design changes of the experimental tire. Even
within one so-called soil type, traction conditions have a great in-
fluence on performance.
Table M shows the force required to pull two types of trailers
across a soil having various cone indexes. The cone index indicates a
combined change of various physical properties in the soil that re-
flects a change in strength. The data indicate that soil has a large
effect on towing force, certainly as large as the difference between the
tracked trailer and the wheeled trailer. For any vehicle, therefore,
the soil condition will determine the traction performance to a much
greater extent than any operational control factor or design factor.
Because of the large influence of soil conditions on traction per-
formance, predicting performance becomes extremely important when
a device is to be operated on a wide range of soil conditions. In
those circumstances (for example, military operations) a means of
predicting performance would provide a means for determining
whether the vehicle will be useful—that is, capable of accomplishing
its purpose. Even further, however, if alternate soil conditions or
paths of travel are available, a suitable path can be selected so that
in effect the vehicle is adapted to the land.
Because of the importance of predicting magnitude of perform-
ance, many attempts have been made to relate performance to soil
conditions. Two of the more successful attempts have been discussed.
As was pointed out, the relations between traction and transport per-
formance must describe the forces in the area of contact. Since this
nature is extremely complex, empirical correlations have been sub-
stituted in lieu of a more fundamental approach. The nature of the
behavior must be considered even in empirical correlations, or the
application of the correlations will be extremely limited. Truly, Sir
Isaac Newton must have seen many apples fall before he was able to
conclude the reason for falling. Similarly, the reasons for traction
and transport performance must be determined if behavior is to be
adequately described. Four aspects of design were outlined in the
introduction to section 7.4. Kesearch concentrating on these four
aspects will give insight into the reasons for traction performance and
eventually lead to a solution of the problem.
are not particularly useful for design because they emphasize pre-
dicting performance, because they are not sufficiently accurate, or
because they are restricted to special situations. Nevertheless, these
equations are examples of empirically developed traction equations.
Soil behavior was shown to be associated with soil physical prop-
erties in sections 3.1 and 6.1. The possibility thus exists of es-
tablishing a correlation between a classification of these properties
and vehicle performance. For a number of years soils have been
classified with respect to their formation and use for crop production
( 4,^6^ JfJf7^ Ji^Iß ). More recently soils have been classified for con-
struction and trafficability purposes ( 10^ 63^ 1^50^ JßJi^ ^95 ). The
U.S. Army Corps of Engineers ( ^9^ ) and the Bureau of Eeclama-
tion ( 450^ 477 ) have unified these classifications to some extent. The
classifications are based on physical properties that are determined
by standardized methods and that indicate certain behavior charac-
teristics. The association of the physical properties and behavior
characteristics is the justification for trying to relate classification
of soil properties to vehicle performance.
Physical properties that have been measured and used as a key
or index to classification include: graduation of particle size, con-
sistency, porosity or void ratio, specific gravity, moisture content,
bulk density, penetration resistance, unconfined compressive strength,
and soluble salts ( 15, 99,18^, 258, ^50 ). With some experience and
judgment one can interpret these indexes in terms of dynamic be-
havior and relate them to vehicle performance.
Texture of soil is one method of classifying soil physical properties
into possible types of behavior reactions. Soil mapping programs
based on texture classification have been made for many soils, and
this information is readily available. One of the most widely used
is the USDA Soil Classification System (fig. 287, A), The Unified
Soil Classification System is shown in figure 287, B, As figure 287
shows, knowledge of a particular soil in one system can be translated
to the other system. Textural classifications are based on the particle
sizes of the mineral constituents of a soil. Moisture content, organic
matter, and structural condition are not reflected in a textural classi-
fication. Organic matter is usually a small fraction of the total
soil volume and remains relatively constant with time. Wlien the
organic matter is a small fraction, its presence is usually ignored.
Highly organic soils are encountered, however, and their physical
properties are often influenced more by organic content than by
mineral content. Mucks and peats are examples of organic soils,
and new classification systems are being developed for them ( 389,
U9).
Several attempts have been made to integrate soil classification and
machine factors into a scheme that would predict performance.
Often these schemes were not represented by an equation, but they
were represented graphically so that a quantitative prediction was
possible. The most widely known scheme utilized the Casagrande
( 63, Í21 ) soil classification to design airfields. The strength of
soil, determined by the California Bearing Ratio test ( 369 ) or from
soil classification data, was correlated with the thickness of pave-
ments and base materials that were required to support aircraft of
422 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
100 i
100
100
SAND (7o)
^
(A)
100
100
SAND (%)
^
(B)
FIGURE 287.—A, USDA Soil Classification System; B, Unified Soil Classifica-
tion System. (Waterways Experiment Station ( W )•)
CL LEAN CLAY 85
SM SILT Y SAND 85
ML SILT 77
OL ORG. SILT 54
PT PEAT, MUCK 45
100
STICKY SOILS
WHEELED VEHICLES
o
o:
O ^30,000 lbs.
FIGURE 289.—Typical relations between the rating cone index and the towing
force required for sleds and vehicles. (Foster, Knight, and Rula, Water-
ways Experiment Station ( 125 ).)
meter and partly to limiting soil conditions to loose sand and wet
saturated clay. These conditions provide the poorest trafficable sit-
uations and so are logical restrictions. As figure 290 shows, more
accurate predictions are obtained than by the scheme shown m figure
288. The added accuracy results because the condition of a soil is
considered and more factors that affect traction are assessed and in-
cluded in the prediction relation.
WHEELED VEHICLES
0.6
• »
0.2
10 20 30 40 50 60
w
c,bd
1500
fiD 1200
900
3
Û.
600
300
CALCULATED RANGE
0-, MEASURED .VALUES
10 20 30 40 50 60 70 80
SLIP (PERCENT)
FIGURE 292.—Calculated and measured pull of a farm tractor in a Michigan
farm sou. (Harrison, 1st Internatl. Conf. Soil-Vehicle Systems Proc. {116).)
farm soil. Other vehicles were also studied and the measurements
and calculations repeated in a dry sand. The difference between
calculated and measured pulls shown in figure 292 is typical of the
difference in performance predicted for other vehicles in sand soil
and is no greater than for any other prediction equation reported in
the literature.
The derived traction equation is the only scheme that attempts to
consider slip. Since moré than the maximum pull is predicted, it is
a more complicated scheme. More calculations and measurements
are required to predict a pull, and the vehicle characterization is more
complete in the derived scheme than the other empirically developed
schemes. A traction equation based on a mechanics has the inherent
means to fully represent all of the factors involved in traction. An
SOIL DYNAMICS IN TILLAGE AND TRACTION 429
8.1 Introduction
In previous chapters, the complete scope of interest of soil dynam-
ics—from descriptions of the soil through its behavior and use—has
been developed. In agriculture, the compactness of soil affects the
soil physical environment for crop production (S, 31, 52, 68, ISO,
m m, 219, 257, 298, 299, 1,13, m, 516 ). Compaction reduces the
soil s permeability to water, so that runoff and erosion may occur and
adequate recharge of ground water is prevented. Compaction re-
duces aeration of the soil, so that metabolic activities of roots are
hampered. Compaction increases the mechanical strength of the
soil, so root growth is impeded. All of these effects may reduce the
quality and quantity of food and fiber grown on the soil.
Direct cause-and-effect relations appear to exist between the use
of machinery and soil compaction, between soil compaction and a
plant root environment, and between a plant root environment and
crop production. These relations are qualitative, and intuition has
led workers to try to directlv correlate the use of machinery with
crop production. Because of the complexity of this machine-plant
system, little progress has been made toward a solution. Less com-
plicated systems must be analyzed to delineate the soil behaviors that
describe and control subsystems wherein soil behavior has direct
effect. When this is accomplished, information will be available to
evaluate and control soil compaction in soil management systems.
Two problems generally arise concerning soil compaction. Soil
may be too compact to be used effectively in crop production. Pre-
vention and alleviation of compactness are the problems associated
with crop production. On the other hand, soil may not be compact
enough to be used effectively for roads, dams, or building supports.
Obtaining maximum compactness with minimum compacting effort is
one of the problems associated with construction work.
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 (sec. 3.1), a change
m state of compactness indicates a probable change in behavior prop-
erties. Hence, regardles of its intended use, the soil is affected by
compaction. General awareness of this fact has stimulated much
soil compaction research. The goal of the research is to solve the
two problems of compaction—to increase or decrease the state of
compactness.
Soil compaction is the action of soil becoming more compact. The
action increases the state of compactness—actions that make soil
430
SOIL DYNAMICS IN TILLAGE AND TRACTION 431
MOISTURE
10 20 30 40 50 60 70
O SHEAR BEGINS
_2.0
E
3 1.6 ^-NORMAL LOAD .118 Kg/cm*
0.6
0.4
T SPECIFIC VOLUME AT SATURATION
SHEAR
0.2
Q.
< _L- -L.
100 300 500 700 900 100 200 300 400
TIME (sec) TIME (min)
mrnirnmiimii —I—
t
I
12.5 K»/ciii*
SANDY LOAM
1.6
1.5 CLAY
ro
E
o
1.4
E
SILT LOAM
t 1.3
CO
z
UJ
Û
1.2
OADING
I.I
L LOADING
1.0
5 10 15 20 30 40 50 60
(^m+ ^m/max)
FIGURE 297.—Compaction relations for several soils. (Vanden Berg, Amer.
Soc. Agr. Engin. ( ^57 ).)
1k
1.9
y"'^ SANDY LOAM
1.8 ■f
1.7 "1
1
"eO
^
E 1.6
o»
>-
1.5 ^ ¥^ SILTY LOAM
^ /
1.4 -/
//
M
-J
3
CD
1.3
f
1.2
u.
0 5 K) 15
DUMBER OF IMRACTS
INSTRUMENTED^
DIAPHRAGM
(PORE PRESSURE)
OIL' FILLED OIL' FILLED X INSTRUMENTED
CONNECTION POROUS STONE DIAPHRAGM
(VERTICAL STRESS)
txJ
DC
^_A»ÊC—^
kl.01,«:
4 lb/In*
IMPACT KNEADING
400 r
<0
(O
bJ KNEADING (300 PSI)
a 200
KNEADING (200 PSI)
ï 100
-KNEADING (100PSI)
10 15 20 25 30
MOISTURE CONTENT (%)
1.90
E
^1.80
É
^1.70 H
O)
g 1-60
Û
§ 1.50 ^
1.40
"i 1 ' ^ 7Z
6 8 10 12 14 16 18 20
MOISTURE CONTENT (7o)
behavior equations for forces of this type, usually the forces are
not identified directly; rather, the cause of the forces is described.
Dryins:, in which the action of soil compaction is termed shrinkage
(fig. 303), is a typical force system of this type. Note that the be-
havior input is expressed as percentage of moisture loss and not ab-
solute moisture content. The data indicate that compaction caused
bv drying can be as great as that caused by mechanical torces.
However, the losses for these soils (20 to 30 percent) require the
input of significant amounts of drying energy. Usually moisture
losses in soils are lower except at an exposed surface. Nevertheless,
natural forces can cause excessive compaction and natural torce-
compaction relations will have to be represented before compaction
can be predicted and controlled.
8.3 Compaction in Tillage and Traction
Soil compaction is important in tillage and traction because of its
effect on the intended use of a soil. The consequences of compaction
are of direct concern rather than compaction per se. These conse-
quences are manifest in terms of specific active and passive behavior
equations that are associated with use of the soil (sec. 6.2). Qualita-
tive data that set limits on the desired state of compactness are
available. Sufficient compactness to establish seed-soil contact and
enhance germination is desirable in the immediate vicinity o± seed.
Excessive compactness, on the other hand, is detrimental to maintain-
ing a good root environment. Excessive compactness also reduces
penetration of water and increases runoff and erosion. The intent
is to control compaction during tillage and traction operations and to
442 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
10 15 20 25 30 35
WATER LOSS (%)
FIGURE 303.—Soil compaction caused by drying: A and B, two clay soils;
C, a silt loam.
DEPTH BELOW
SOIL SURFACE
(INCHES)
0
I 2 3 4 5 6
DISTANCE - (FT.)
o 25
o 35
15 10 5 0 5 10 15 15 to 5 0 5 10
FIGURE 305.—Mean normal stress under a track and a wheeL (Reaves and
Cooper, Agr. Engin. ( SJfJf ).)
8 O.SOr
0.25
X
lu
t 0.00
8 »00 1.20 \30 L40 1.50 1.60 1.70 1.80
a:
BULK DENSITY (gm/cm*)
E
u 800
t ñ
600
Z Ü
400
^ hi
200
100 200
ANGULAR STRAIN (*»)
100 r
g
11 % MOISTURE CONTENT
<
z
o
¡5
OC
CO
O
r=- >
ÛL
cc
o
CO
LU
Û
O
h-
O
<
o
>-
" i
FRACTIONAL COMPLETE
COMPLETENESS OF REPRESENTATION
FIGURE 309.—Definition of soil-machine systems.
SOIL DYNAMICS IN TILLAGE AND TRACTION 449
TRANSMITTED
SUPPLIES OF AIR, HEAT,
AND WATER
MECHANICAL SUPPORT,
GROWING SPACE, LOW
MECHANICAL IMPEDANCE
PREDOMINANT AREAS OF INTEREST ■■■■ SOIL GENESIS, PHYSICS ■■■ SOIL-MACHINE DYNAMICS
FOR EACH DISCIPLINE • • • • SOIL PHYSICS !•■• SOIL PLANT DYNAMICS
how soil dynamics is associated with plant growth. Figure 310 also
shows that soil dynamics is a newly defined area or discipline con-
cerning the dynamic action of the soil in tillage. The bulk of past
work on soil mechanics has been associated with the needs of earth-
work construction and foundations for large structures. As such,
emphasis was placed on phenomena like the slow settling of soil under
buildings—phenomena that are very different from the dynamic
phenomena involved in vehicle mobility and tillage. Soil tilth also
may evolve from genetic factors, but the active behavior of the soil
under these forces is generally very slow.
A second newly defined area is soil-plant dynamics. As m soil
genesis, the active behavior of the soil is so slow that the force re-
quired to accelerate the soil is essentially zero during movement by
plant roots. Not enough work has been done to define this area, but
the scope and importance of the work is sufficient to warrant future
development. . i • i i i •
Soil physics as viewed here is generally associated with the physi-
cal transmission activities of the soil in which soil is passive. Prog-
452 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
ress in knowledge of soil physics has been largely in terms of trans-
mission of soil moisture and soil air. Thus, if an area is recognized
where the soil is active rather than passive, specialized courses can
be taught and additional research can be directed toward obtaining
basic knowledge concerning this important soil behavior.
One final word can be said concerning soil dynamics. The farmer
has effectively integrated soil relations in his system of interest by
physical operations that are qualitative rather than quantitative;
integration will not be as easy on a technical level.
Kesearch generally develops information that separates into di-
vergent areas of greater detail rather than into simpler and more
complete bodies of knowledge. This handbook provides a frame-
work m which the independent efforts of individual workers dealing
with soil dynamics can be placed into perspective. Thus, even
though some workers proceed into divergent areas in great detail,
the relation and significance of the work can be realized and utilized
by others. It is this realization that will provide common areas of
interest for coordinating efforts in accepting and utilizing data from
other sources. While some workers are developing narrow areas of
application in detail, others may be unifying the problems of broader
scope without mutual delay or interference. This unifying aspect
of the organization of soil relations needs to be stressed in imple-
menting all work in soil dynamics. Soil dynamics recognized as a
discipline can serve this function.
9.3 Soil Dynamics and Tillage
Tillage is defined as the physical manipulation of the soil; this
definition is broader than the classical concept since it includes such
operations as land forming, earth moving, and soil conditioning.
These operations may be performed with the same tillage tools and
in the same manner as classical tillage operations even though crop
production is not contemplated. The fundamentals in the actions
of various soil-machine systems are the same, and soil dynamics de-
scribes these actions. Hence, there is no need to restrict the field
of application of soil dynamics to any specific purpose for the actions.
The dynamic actions in soil-tillage tool systems are important be-
cause of the power used in tillage. McKibben ( 268 ) has calculated
that for agriculture alone, 250 billion tons of soil are stirred or turned
each year in the United States. This volume of soil represents a
ridge 100 feet high and 1 mile wide extending from San Francisco
to New York. To plow this ridge once requires 500 million gallons
of gasoline. Even small savings of liquid fuels can be significant on
a national basis.
Energy savings may be influenced by factors other than those that
govern the operating efficiency of a tillage tool itself. The efficiency
of a tool operation can be increased only to the extent that the tillage
objectives represent minimum energy requirements or expenditures.
These objectives are determined by the requirements of the broad
system in which the soil-tool system operates. A reflection on the
tillage objectives that have been advocated in the past leads to the
conclusion that a new philosophy of tillage is required. An acre
of soil 6 inches deep weighs 2 million pounds. Plowing this acre
SOIL DYNAMICS IN TILLAGE AND TRACTION 453
trips over the soil does not necessarily constitute minimum tillage.
A clear-cut example of minimum tillage is land forming. Com-
puters have been used to calculate the minimum soil that needs to be
moved to produce acceptable grades. By varying grades within
prescribed limits, a smaller volume of soil needs to be handled.
A sequence of tillage operations to achieve a desired soil condition
is not a sacrosanct practice; rather, it is a necessity dictated by avail-
able tillage tools. The apparent philosophy guiding tillage tool de-
sign today is that each type of tool shall accomplish a specific action
and that this action will be universally applicable. Such a philoso-
phy is absurd since the soils and soil conditions that must be manip-
ulated vary greatly. The tillage equations illustrate that the result-
ing soil condition is a function of the initial soil condition. For a
constant shape and constant manner of movement (a universal tool),
the resulting condition is determined by the initial soil condition ; it
is not controlled by the tillage tool. If the resulting condition is not
suitable, the tiller selects a different tool. Again, the resulting soil
condition will be primarily determined by the condition before the
operation, since shape and manner of movement are again constant.
Because today's tiller has only limited control over the shape and
manner of movement of a tool, he is forced to do sequence opera-
tions. Only by selecting different tools in a sequence can he change
either the shape or the manner of movement or both, so as to produce
the desired final soil condition.
The present necessity of sequence operations can be avoided by
two approaches. ^ Eliminating a sequence of operations means, in
effect, that the soil will be manipulated to the desired condition in a
one-step operation. Intuitively, proceeding from initial condition
to final condition as directly as possible should require the least
energy. Since many initial conditions will be encountered and many
final conditions will be desired, many tool shapes and manners of
movement will be required. For a specific situation, the tiller must
have some means of selecting the combination of shapes and manner
of movement that will accomplish the manipulation. He must have
either a large number of shapes or a shape that he can vary and
control. He must also have either a large number of mechanisms
that will vary the manner of movement or one mechanism that he
can vary and control. The two approaches for eliminating sequence
operations thus are (1) a large number of tillage machines, each
suitable for one specific circumstance; (2) a tillage machine that
the tiller can adjust for a specific circumstance.
A combination of the two possibilities may prove to be the prac-
tical solution. To give the tiller control over soil manipulation, till-
age machines will have to become more complex ; simultaneously, the
tiller will have to become a proficient technician, capable of properly
using these machines.
Other concepts of controlling soil conditions are possible. Little
attention has been given to rocks and stones other than to remove
them from the soil. With the advent of inexpensive yet powerful
sources of energy, undesirable soil conditions may be corrected by a
mechanical breakdown of particles in the soil large enough to hinder
operations. On the other hand, artificial rocks could probably be
SOIL DYNAMICS IN TILLAGE AND TRACTION 455
(39) BERRY, M. O.
1948. COULTER TESTS. U.S. Tillage Mach. Lab. (Auburn, Ala.)
Ann Rpt.
(40) BERTELSEN, W. R.
1960. THE AEROMOBILE : A PERIPHERAL JET VEHICLE. AffF. Engin 41 *
290-292, illus.
(41)
1960. SOMETHING BRAND NEW IN PLOWS THE AEROPLOW. DeS
Moines Register. Aug. 21, Sect. H, p. 17-H.
(42) BiGSBY, F. W.
1961. EFFECT OF AN AIR FILM ON SOIL TO TILLAGE SURFACE FRICTION.
Ph.D. thesis. Iowa State Univ., 85 pp., illus.
(43) BiKERMAN, J. J.
1958. SURFACE CHEMISTRY. Ed. 2, 501 pp., illus. New York
(44) BJORCK, G.
1958. STUDIES ON THE DRAUGHT FORCE OF HORSES. Acta Agr.
Scand. Sup. 4, 109 pp., illus.
(45) BOA, W.
1958. DEVELOPMENT OF N.I.A.E. DITCH CLEANER. Jour. AsT Engin
Res. 3: 17-26, illus.
(46) and WHYTE, P.
1960. THE DESIGN OF A PLOUGH FOR LAYING UNDERGROUND CABLES.
Jour. Agr. Engin. Res. 5: 135-140, illus.
(47) BODMAN, G. B.
1949. METHODS OF MEASURING SOIL CONSISTENCY. Soll Sci 68* 37-
56, illus.
(48) and RUBIN, J.
1948. SOIL PUDDLING. Soil Sci. Soc. Amer. Proc. 13: 27-36 illus
(49) BONMARTINI, G. '
1961. THE WHEEL: ROADS—THE TRACKS: NATURAL SOILS. 272 pp.
illus. Rome. '
(50) BowDEN, F. P., and TABER, D.
1950. THE FRICTION AND LUBRICATION OF SOLIDS. 337 pp., illuS.
Oxford.
(51) BowEN, H. D.
1960. SOME PHYSICAL IMPEDANCE AND AERATION EFFECTS ON PLANTED
,^^, ^ SEEDS. Amer. Soc. Agr. Engin. Paper 60-626, 19 pp., illus.
(52) BoYD, M. M.
1959. COMPACTION AND SLIPPAGE EFFECTS OF FORAGE HARVESTING
MACHINERY. Amer. Soc. Agr. Engin. Paper 59-102, 7 pp
illus.
(53) BROWNING, G. M.
1950. PRINCIPLES OF SOIL PHYSICS IN RELATION TO TILLAGE Affr
Engin. 31: 341-344.
(54) BRUCE, R. R.
1955. AN INSTRUMENT FOR THE DETERMINATION OF SOIL COMPACTI-
,^^, BiLiTY. Soil Sci. Soc. Amer. Proc. 19: 253-257, illus.
(55) BRU WER, J. J.
1962. THE INSTANTANEOUS TRACTIVE EFFICIENCY OF A 12-38 PNEU-
MATic TRACTOR TIRE. Amer. Soc. Agr. Engin. Trans. 5: 114.
(56) BUSCH, C. D.
1958. LOW COST SUBSURFACE DRAINAGE. Agr. Engin. 39: 92-93
97, 103, illus. * '
(57)
1958. MECHANICAL MOUSE AIDS RESEARCH IN SUBSURFACE DRAINAGE
Agr. Engin. 39: 292-293, illus.
(58) CAREW, N.
/^^^ •^^^^* ®^^^ COMPACTION. Sugar Jour. 18: 35-38, illus.
(59) CARLETON, W. M., and MARTIN, J. W.
1945. SHARPENING AND HARDSURFACING PLOW AND LISTER SHARES
Kans. Engin. Expt. Sta. Bui. 44, 40 pp., illus
(60) CARLSON, E. C.
1961. PLOWS AND COMPUTERS. Agr. Engin. 42: 292-295, 307, illus.
SOIL DYNAMICS IN TILLAGE AND TRACTION 463
CABNES, A.
1934. SOIL CRUSTS METHODS OF STUDY, THEIR STRENGTH, AND A
METHOD OF OVERCOMING THEIR INJURY TO COTTON STAND. Agr.
Engin. 15: 167-171, illus.
CARPENTER, F. G., and DEITZ, V. R.
1951. GLASS SPHERES FOR THE MEASUREMENT OF THE EFFECTIVE
OPENING OF TESTING SIEVES. [U.S.] Nat'l. Bur. Standards
Jour. Res.. 47 : 139-147, illus.
CASAGRANDE, A.
1947. CLASSIFICATON AND IDENTIFICATION OF SOILS. AMER. SOC.
Civ. Engin. Proc. 73: 783-810, illus.
DAVIES, D. L.
1954. THE DESIGN AND ANALYSIS OF INDUSTRIAL EXPERIMENTS. 636
pp., illus. London and Edinburgh.
DAVIS, W. M.
1961. IMPLEMENT REQUIREMENTS IN RELATION TO TRACTOR DESIGN.
Agr. Engin. 42: 478-483, illus.
DAY, P. R..
1950. PHYSICAL BASIS OF PARTICLE SIZE ANALYSIS BY THE HYDRO-
METER METHOD. Soil Sei. 70: 363-374, illus.
and HOLMGREN, G.
1952. MICROSCOPIC CHANGES IN SOIL STRUCTURE DURING COMPRESSION.
Soil Sei. Soe. Amer. Proc. 16: 73-77, illus.
DELEENHEER, L., and DEBOODT, M.
1958. DETERMINATION OF AGGREGATE STABILITY BY THE CHANGE IN
MEAN WEIGHT DIAMETER, Intomatl. Syuipos. OU Soil Struc-
ture Proe. (Ghent, Belgium) pp. 290-300, illus.
DICKSON, W. J.
1962. GROUND VEHICLE MOBILITY ON SOFT TERRAIN. AMER. SOC. CIV.
Engin. Proc. 88 (SM4, pt. 1) : 69-83, illus.
DiNGLINGER, E.
1932. THE CUTTING OF SAND. Engineering 134: 116-118.
DOMSCH, M.
1955. [PROBLEMS OF SOIL CULTIVATION.] 140 pp., illus. Berlin.
DONER, R. D.
1936. A THEORY OF ARCH ACTION IN GRANULAR MEDIA. Agr. Eugiu.
17: 299-304, illus.
and NICHOLS, M. L.
1934. THE DYNAMIC PROPERTIES OF SOIL. V. DYNAMICS OF SOIL ON
PLOW MOLDBOARD SURFACES RELATING TO SCOURING. Agr.
Engin. 15: 9-13, illus.
DUBROVSKII, A. A.
1956. [INFLUENCE OF VIBRATING THE TOOLS OF CULTIVATION IMPLE-
MENTS UPON DRAFT RESISTANCE.] Vsesoiuzz. Akad. Scl' skok-
hozhaistvenny k h Nauk. Zeml. Mekh. Shorn. Trudov.
(Leningrad) 3: 182-185, illus. [Nati. Inst. Agr. Engin.,
Eng. Translation 51.]
Dupuis, H.
1959. EFFECT OF TRACTOR OPERATION ON HUMAN STRESS. Agr. EugiU.
40: 510-519, 525, illus.
ECKMAN, D. P.
1960. SYSTEMS : RESEARCH AND DESIGN. 310 pp., illus. New York.
EGGEN MÜLLER, A.
1958. [OSCILLATING IMPLEMENTS : KINEMATICS AND EXPERIMENTS
WITH MODELS.] Grundlagen der Landtechnik 10: 55-69, illus.
K
466 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(211)
1957. [CHARACTERISTICS OF SOIL RESISTANCE OF VARIOUS SHAPE
MOLDBOARD PLOWS.] Osaka [Japan] Univ. Bui. Ser. B 7:
65-75, illus.
(212)
1957. [EXPERIMENTAL STUDIES OF PLOW SHAPES.] SOC. Agr. Mach.
Jour., (Japan) 19: 97-100, illus.
(213)
1958. DYNAMIC BEHAVIOR OF SOIL. II. HIGH SPEED TRIAXIAL COM-
PRESSION TEST. Soc. Agr. Mach. Jour. (Japan) 20: 101-
103, illus.
(214) KEEN, B. A.
1931. THE PHYSICAL PROPERTIES OF THE SOIL. 380 pp., lUUS. NCW
York.
(215) and HAINES, W. B.
1925. STUDIES IN SOIL CULTIVATION. Jour. Agr. Sei. 15: 375-406,
illus.
(216) KERESELIDZE, S. Y., KHUKHUNI, T. V., and SHKOLNIK, E. B.
1959. [INVESTIGATIONS INTO THE OPERATION OF THE AUTOMATIC
STABILIZER OF THE USG-12A HILLSIDE TRACTOR.] TraktOry Í
Sel'khozmashiny (Moscow) 3: 4, illus. [Jour. Agr. Engin.
Res. 4: 267-273, Eng. Translation.]
(217) KHAMIDOV, A.
1961. [USE OF GAMMA-RAYS IN THE STUDY OF THE EFFECTS OF
WHEELS ON THE SOIL.] Trudy VIM 28: 94-107, illus. [Jour.
Agr. Engin. Res. 6: 147-152, Eng. Translation.]
(218) KiRKHAM, D., DEBOODT, M. F., and DELEENHEER, L.
1959. MODULUS OF RUPTURE DETERMINATION ON UNDISTURBED SOIL
CORE SAMPLES. Soil Scl. 87: 141-144, illus.
(219) KLIEFOTH, F.
1954. [STRUCTURE DAMAGE THROUGH THE SLIDING OF WHEELS.]
Landtechnik 9: 122-123, illus.
(220) KNIGHT, S. J. and FREITAG, D. R.
1962. MEASUREMENT OF SOIL TRAFFICABILITY CHARACTERISTICS.
Amer. Soc. Agr. Engin. Trans., 5: 121-124, 132, illus.
(221) KNOX, E. G.
1957. FRAGIPAN HORIZONS IN NEW YORK SOILS. III. THE BASIS
OF RIGIDITY. Soil Sci. Soc. Amer. Proc. 21: 326-330, illus.
(222) KOLCHIN, N. N.
1957. [COMBINED PNEUMATIC AND MECHANICAL SEPARATION OF POTATO
TUBERS FROM CLODS.] Sel'khozmashina 3 : 19-21, illus. [Jour.
Agr. ^ngin. Res. 2: 238-240, Eng. Translation.]
(223) KONDNER, R. L.
1958. A NON-DIMENSIONAL APPROACH TO THE VIBRATORY CUTTING,
COMPACTION AND PENETRATION OF SOILS. Tech. Rpt. 8 by the
Johns Hopkins Univ. to U.S. Army Corps of Engin., Water-
ways Expt. Sta., Vicksburg, Miss., 183 pp., illus.
(224)
1959. THE VIBRATORY CUTTING, COMPACTION AND PENETRATION OF
SOILS. Tech. Rpt. 7 (Part II), by the Johns Hopkins Univ.
to U.S. Army Corps of Engin., Waterways Expt. Sta.,
Vicksburg, Miss., 51 pp., illus.
(225)
1962. A PENETROMETER STUDY OF THE IN SITU STRENGTH OF CLAYS.
Mater. Res. and Standards 2: 193-195, illus.
(226) and AYRE, R. S.
1959. STUDY OF VIBRATORY CUTTING, PENETRATION AND COMPACTION
OF SOILS. Tech. Rpt. 6 by the Johns Hopkins Univ. to U.S.
Army Corps of Engin., Waterways Expt. Sta., Vicksburg,
Miss., 7 pp., illus.
(227) AYRE, R. S., and CHAE, Y. S.
1958. LABORATORY INVESTIGATION OF THE VIBRATORY CUTTING AND
PENETRATION OF SOILS (part 1). Tcch. Rpt. 5 by the Johns
Hopkins Univ. to U.S. Army Corps of Engin., Waterways
Expt. Sta., Vicksburg, Miss., 12 pp., illus.
472 AGRICULTURE HANDBOOK 816, U.S. DEPT. OF AGRICULTURE
(228) KOOPMAN, N. P.
1957. [RESEARCH ABOUT THE WEAR OF PLOW SHARES.] Rpt. Agr.
State Univ., Dept. Agr. Engin., Wageningen, Netherlands,
illus.
(229) KORAYEM, A. Y.
1961. FRicTiONAL PROPERTIES OF ARTIFICIAL SOILS. Unpublished M.
S. Thesis, Univ. of Calif., illus.
(230) KosTRiTSYN, A. K.
1956. [CUTTING OF A COHESIVE MEDIUM WITH KNIVES AND CONES.]
Vsesoiuzz. Akad. Sel'skokhoziaistvennykh Nauk. Zeml. Mekh.
Shorn. Trudov. (Leningrad) 3: 247-290, illus. [Nati. Inst.
Agr. Engin., Eng. Translation 58.]
(231) KRUTIKOV, N. P., SMIRNOV, I. I., STSCHERBAKOV, K. F., and POPOV, I. F.
1955. THEORY, CALCULATION AND DESIGN OF FARM MACHINERY. V, I.
MACHINERY FOR SOIL CULTIVITION, SOWING, AND PLANT CUL-
TIVATION. 690 pp., illus. Berlin.
(232) KUHN, G.
1954. [SHAPE OF BULLDOZER BLADES TO ATTAIN THE SMALLEST FILLING
RESISTANCE.] Ver. Deut. Ingen. Ztschr. 96: 373-377, illus.
(233)
1954. [SHAPE OF BULLDOZER BLADES TO ATTAIN THE SMALLEST FILLING
RESISTANCE.] Ver. Deut. Ingen. Ztschr. 96: 982-986, illus.
(234) KUMMER, F. A.
1939. THE DYNAMIC PROPERTIES OF SOIL. VIII. THE EFFECT OF
CERTAIN EXPERIMENTAL PLOW SHAPES AND MATERIALS ON
SCOURING IN HEAVY CLAY SOILS. Agr. Eugiu. 20: 111-114,
illus.
(235) and NICHOLS, M. L.
1938. THE DYNAMIC PROPERTIES OF SOIL, VII. A STUDY OF THE
NATURE OF PHYSICAL FORCES GOVERNING THE ADHESION BE-
TWEEN SOIL AND METAL SURFACES. Agr. Eugiu. 19: 73-78,
illus.
(236) KYNAZEV, A. A.
1957. [DRAUGHT RESISTANCE OF CHISEL PLOUGHS.] Sel'khozmash-
ina 2: 7-9, illus. (Nati. Inst. Agr. Engin., Eng. Transla-
tion 89.)
(237) LAL, R.
1959. MEASUREMENT OF FORCES ON MOUNTED IMPLEMENTS. Amer
Soc. Agr. Engin. Trans. 2: 109-111, illus.
(238) LAMBE, T. W.
1951. SOIL TESTING FOR ENGINEERS. 165 pp., illus. New York
(239) LAMBE, T. W.
1958. THE STRUCTURE OF COMPACTED CLAY. Ainer. Soc. Civ. Engin
Proc. 84 : Paper 1654, 34 pp., illus.
(240)
1958. THE ENGINEERING BEHAVIOR OF COMPACTED CLAY. Amer. SoC.
Civ. Engin. Proc. 84 : Paper 1655, 34 pp., illus.
(241) LAND LOCOMOTION LABORATORY, DETROIT ARSENAL, WARREN, MICH.
1956. PRELIMINARY STUDY OF SNOW VALUES RELATED TO ' VEHICLE
PERFORMANCE. Tech. Memo M02, 27 pp., illus.
(242)
1957. AN INVESTIGATION OF GUN ANCHORING SPADES UNDER THE AC-
TION OF IMPACT LOADS. Rpt. 19, 33 pp., illus.
(243)
1957. ARTIFICIAL SOILS FOR LABORATORY STUDIES IN LAND LOCOMOTION
Rpt. 20, 29 pp., illus.
(244)
1957. STUDY OF SNOW VALUES RELATED TO VEHICLE PERFORMANCE
Rpt. 23, 31 pp., illus.
(245)
1958. A SOIL VALUE SYSTEM FOR LAND LOCOMOTION MECHANICS ReS
Rpt. 5, 94 pp., illus.
(246)
1960. EVALUATION OF CONDUAL TIRE MODEL. Rpt. 60, 28 pp., illuS.
SOIL DYNAMICS IN TILLAGE AND TRACTION 473
(247)
1960. THE MECHANICS OF WALKING VEHICLES. Rpt. 71, 67 pp.,
illus.
(248)
1961. BEVAMETER 100. A NEW TYPE OF FIELD APPARATUS FOR MEAS-
URING LOCOMOTIVE STRESS-STRAIN RELATIONSHIPS IN SOILS.
1st Internatl. Conf. Mech. Soil-Vehicle Systems Proc, (Tu-
rin), pp. 331-361, illus.
(249)
1961. THE ANALYTICAL DETERMINATION OF DRAWBAR PULL AS A
FUNCTION OF SLIP FOR TRACKED VEHICLES IN DEFORMARLE SOILS.
1st Internatl. Conf. Mech. Soil-Vehicle Systems, Proc. (Tu-
rin), pp. 707-736, illus.
(250) LANGE, H.
1957. [ABILITY OF PNEUMATIC TIRES TO ABSORB THE LATERAL FORCES
ACTING ON AGRICULTURAL MACHINES WHEN TRAVELING ACROSS
THE SLOPE.] Grundlagen der Landtechnik 9: 109-112, illus.
(251) LASZLO, K.
1951. [EFFECT OF TILLAGE MACHINES ON SOIL STRUCTURE.] Expt.
Inst. for Agr. Mach. ybk. (Budapest) v. II, p. 175-208, illus.
(252) LECHNER, F. G. and MCCOLLY, H. F.
1959. ABRASIVE-WEAR RESISTANCE OF HARD-FACING MATERIALS USED
ON AGRICULTURAL TILLAGE TOOLS. Amcr. SOC. Agr. EugiU.
Trans. 2: 55-57, illus.
(253) LEWIS, W. A.
1954. FURTHER STUDIES IN THE COMPACTION OF SOIL AND THE PER-
FORMANCE OF COMPACTION PLANT. [Gt. Brit.] Dept. Sei.
and Indus. Res. Road Res. Lab., Tech. Paper 33, illus.
(254)
1957. A STUDY OF SOME OF THE FACTORS LIKELY TO AFFECT THE
PERFORMANCE OF IMPACT COMPACTORS ON SOIL. 4th Internatl.
Conf. Soil Mech. and Found. Engin. Proc. (London) 2: 145-
150, illus.
(255) LoRENZEN, C, JR.
1950. THE DEVELOPMENT OF A MECHANICAL ONION HARVESTER. Agr.
Engin. 31: 13-15, illus.
(256) LOVE, A. E. H.
1929. THE STRESS PRODUCED IN A SEMI-INFINITE BODY BY PRESSURE
ON PART OF THE BOUNDARY. Roy. Soc. Loudon, Phil. Trans.,
Ser. A, 228: 277^20.
(257) LULL, H. W.
1959. SOIL COMPACTION ON FOREST AND RANGE LANDS. U.S. Dept.
Agr. Misc. Pub. 768, 33 pp., illus.
(258) LUTZ, J. F.
1947. APPARATUS FOR COLLECTING UNDISTURBED SOIL SAMPLES. SOIL
Sei. 64: 399-401, illus.
(259) LYSAGHT, V. E.
1947. THE KNOOP INDENTER AS APPLIED TO TESTING NON-METALLIC
MATERIALS RANGING FROM PLASTICS TO DIAMONDS. Amer. SoC.
Testing Mater. Bui. 138, 39 pp., illus.
(260) MACLEAN, D. J., and ROLFE, D. W.
1946. A LABORATORY INVESTIGATION OF ELECTRO-OSMOSIS IN SOILS.
Philos. Mag. 37 : 863-873, illus.
(261) MCCLELLAND, J. H.
1956. INSTRUMENT FOR MEASURING SOIL CONDITION. Agr. Engin.
37: 480-481, illus.
(262) MCCREERY, W. F.
1959. EFFECT OF DESIGN FACTORS OF DISKS ON SOIL REACTIONS. AMCR.
Soc. Agr. Engin. Paper 59-622, 6 pp., illus.
(263) and NICHOLS, M. L.
1956. THE GEOMETRY OF DISKS AND SOIL RELATIONSHIPS. Agr. Eugiu.
37: 808-812, 820, illus.
(264) MCFARLANE, J. S., and TABOR, D.
1950. ADHESION OF SOLIDS AND THE EFFECT OF SURFACE FILMS.
Roy. Soc. London, Proc, S©r. A, 202: 224-243, illus.
474 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(265) and TABOR, D.
1950. RELATION BETWEEN FRICTION AND ADHESION. Roy. SOC. Lon-
don, Proc, Ser. A, 202: 244-253, illus.
(266) McKiBBEN, E. G.
1926. THE SOIL DYNAMICS PROBLEM. Agr. Engin. 7: 412-413, illus.
(267)
1938. SOME KINEMATIC AND DYNAMIC STUDIES OF RIGID TRANSPORT
WHEELS FOR AGRICULTURAL EQUIPMENT. lOWa Agr. Expt. Sta.
Res. Bui. 231, pp. 321-390, illus.
(268)
1959. ENGINEERING IN AGRICULTURE. Agr. Engin. 40: 385, 412,
414, illus.
(269) and BERRY, M. O.
1952. THE VALUE OF REPLICATIONS IN RESEARCH. Agr. Eugiu. 33:
792, 798, illus.
(270) and DAVIDSON, J. B.
1939. TRANSPORT WHEELS FOR AGRICULTURAL MACHINES. II. ROLL-
ING RESISTANCE OF INDIVIDUAL WHEELS. Agr. Engin. 20*
469-473, illus.
(271) and DAVIDSON, J. B.
1940. TRANSPORT WHEELS FOR AGRICULTURAL MACHINES. III. EFFECT
OP INFLATION PRESSURE ON THE ROLLING RESISTANCE OF PNEU-
MATIC IMPLEMENT TIRES. Agr. Engin. 21: 25-26, illus.
(272) and DAVIDSON, J. B.
1940. TRANSPORT WHEELS FOR AGRICULTURAL MACHINES. IV. EFFECT
OF OUTSIDE AND CROSS-SECTIONAL DIAMETERS ON THE ROLLING
RESISTAT^CE OP PNEUMATIC IMPLEMENT TIRES. Agr. Eugiu. 21 :
57-58, illus.
(273) and DAVIDSON, J. B.
1940. TRANSPORT WHEELS FOR AGRICULTURAL MACHINERY. V. EFFECT
OF WHEEL ARRANGEMENT ON ROLLING RESISTANCE. Agr. Euffin.
21: 95-96, illus.
(274) and DAVIDSON, J. B.
1940. TRANSPORT WHEELS FOR AGRICULTURAL MACHINES. VI. EF-
FECTS OF STEEL WHEEL RIM SHAPE AND PNEUMATIC TIRE-
TREAD DESIGN ON ROLLING RESISTANCE. Agr. Engin. 21 : 139-
140, illus.
(275) and DAVIDSON, J. B.
1940. TRANSPORT WHEELS FOR AGRICULTURAL MACHINES. IX. EFFEC-
TIVE RADIUS AND SLIPPAGE. Agr. Eugiu. 21: 275-276. illus
(276) and GREEN, R. L.
1940. TRANSPORT WHEELS FOR AGRICULTURAL MACHINERY. VII.
RELATIVE EFFECTS OF STEEL WHEELS AND PNEUMATIC TIRES ON
AGRICULTURAL SOILS. Agr. Engin. 21: 183-185, illus.
(277) and HULL, D. O.
1940. TRANSPORT WHEELS FOR AGRICULTURAL MACHINES. VIII. SOIL
PENETRATION TESTS AS A MEANS OF PREDICTING ROLLING RE-
SISTANCE. Agr. Engin. 21: 231-234, illus.
(278) and REED, I. F.
1952. THE INFLUENCE OF SPEED ON THE PERFORMANCE CHARACTER-
ISTICS OF IMPLEMENTS. Paper, SAE Nati. Tractor Mtg.
Milwaukee, Wis., 5 pp., illus. '
(279) and THOMPSON, H. J.
1939. TRANSPORT WHEELS FOR AGRICULTURAL MACHINES. I. COM-
PARATIVE PERFORMANCE OF STEEL WHEELS AND PNEUMATIC
TIRES ON TWO MANURE SPREADERS OF THE SAME MODEL. AsT
Engin. 20: 419^22, illus. *
(280) McMuRDiE, J. L.
1963. SOME CHARACTERISTICS OF THE SOIL DEFORMATION PROCESS
,^^_ Soil Sei. Soc. Amer. Proc. 27: 251-254, illus.
(281) and DAY, P. R.
1958. COMPRESSION OF SOIL BY ISOTROPIC STRESS. Soil Sci SoC
Amer. Proc. 22: 18-21, illus.
(282) and DAY, P. R.
1960. SLOW TESTS UNDER SOIL MOISTURE SUCTION. Soil Sci Soc
Amer. Proc. 24: 441-444, illus.
SOIL DYNAMICS IN TILLAGE AND TRACTION 475
(283) MCRAE, J. L.
1950. A STUDY OF STRESSES IN SOIL DURING LABORATORY COMPACTION.
[Unpublished.] Dept. of Civil Engin., Northwestern Univ.,
14 pp., illus.
(284) MAACK, O.
1957. [THE MECHANICAL SEPARATION OF POTATOES AND STONES.]
Landtechnische. (Munich) Forsch. 7: 71-78, illus. [Nati.
Inst. Agr. Engin., Eng. Translation 35.]
(285)
1957. [THE SEPARATION OF POTATOES AND STONES ON THE BASIS OF
RESILIENCE AND ROLLING RESISTANCE,] Landtcchnishc. Forsch.
(Munich) 7: 106-110, illus. [Nati. Inst. Agr. Engin., Eng.
Translation 36.]
(286) MACKSON, C. J.
1962. THE EFFECT OF ELECTRO-OSMOSIS ON SOIL TO STEEL SLIDING
FRICTION AS INFLUENCED BY SPEED, VOLTAGE AND SOIL MOIS-
TURE. Amer. Soc. Agr. Engin. Paper 62-650, 8 pp., illus.
(287) MANLEY, P. L.
1960. AN INVESTIGATION OF MECHANICAL AND PHYSICAL FACTORS
INFLUENCING THE STABILITY OF MOLE DRAINS UNDER LOAD.
M. S. Thesis, Cornell Univ., illus.
(288)
1960. PLASTIC DRAIN LINERS EVALUATED BY PNEUMATIC TEST CHAM-
BER. Agr. Engin. 41: 824, illus.
(289) MARSHAK, A. L.
1956. [SURFACE SHAPE OF PNEUMATIC TIRES IN CONTACT WITH THE
SOIL.] Sel'khozmashina 3: 22-24, illus. [Off. Tech. Serv.,
Translation 60-21084.]
(290)
1957. MACHINES.]
[ROLLING RESISTANCE OF PNEUMATIC TIRES ON FARM
Sel'khozmashina 1: 13-16, illus.
(291) MARSHALL, T. J.
1959. RELATIONS BETWEEN WATER AND SOIL. Commonwcalth Bur.
Soils (Harpenden, England), Tech. Comm. 50, 91 pp., illus.
(292) and QUIRK, J. P.
1950. STABILITY OF STRUCTURAL AGGREGATES OF DRY SOIL. Austral.
Jour. Agr. Res. 1: 226-275, illus.
(293) MARTINI, Z.
1953. [STUDIES ON SOIL MOVEMENT DURING THE PLOWING OF SLOPES.]
Rocz. Nauk Roln. 66-C(2) : 97-123.
(294) MAUER, H. C.
1933. A METHOD OF MEASURING THE PRESSURE OF THE SOIL ON PLOWS
OR OTHER TILLAGE IMPLEMENTS. M. S. ThesiS, Aubum UuiV.,
13 pp., illus.
(295) MAYAUSKAS, I. S.
1958. INVESTIGATION OF THE PRESSURE DISTRIBUTION ON THE SURFACE
OF A PLOW SHARE IN WORK. Traktory i Sel'khozmashiny
11: 23- illus. [Jour. Agr. Engin. Res. 4: 186-190. 1959,
Eng. Translation.]
(296) MECH, S. J., and FREE, G. R.
1942.MOVEMENT OF SOIL DURING TILLAGE OPERATIONS. Agr. EugiU.
23: 379-382, illus.
(297) MENZEL, R. G., ROBERTS, H., JR., and JAMES, P. E.
1961. REMOVAL OF RADIOACTIVE FALLOUT FROM FARM LAND. PrOg.
Rpt. II. Agr. Engin. 42: 69S-699, illus.
(298) MEREDITH, H. L. and PATRICK, W. H., JR.
1961. EFFECTS OF SOIL COMPACTION ON SUBSOIL ROOT PENETRATION
AND PHYSICAL PROPERTIES OF THREE SOILS IN LOUISIANA.
Agron. Jour. 53: 163-167, illus.
(299) MILLER, S. A., and MAZURAK, A. P.
1958. RELATIONSHIPS OF PARTICLE AND PORE SIZES TO THE GROWTH
OF SUNFLOWERS. Soil Sci. Soc. Amcr. Proc. 22: 275-278, illus.
(300) MiscENKO, N. F.
1935. METHODS OF FIXATION AND POROSITY DETERMINATION IN THE
STUDY OF SOIL MECHANICS. Agr. Eugiu. 16: 23-29, illus.
476 AGRICULTURE HANDBOOK 316, US. DEPT. OF AGRICULTURE
(301) MITCHELL, N. B.
1961. THE INDIRECT TENSION TEST FOB CONCRETE, MATERIALS, RE-
SEARCH, AND STANDARDS. Mater. Res. and Standards 1:
780-788, illus.
(302) MOHSENIN, N., WoMOCHEL, H. L., HARVEY, D. J., and CARLETON, W. M.
1956. WEAR TESTS OF PLOWSHARE MATERIALS. Agr. Engin. 37: 816-
820, illus.
(303) MÖLLER, R.
1959. DRAFT REQUIREMENTS AND WORKING EFFICIENCY OF RIGID AND
SPRING CULTIVATOR TINES. Grundlagen der Landtechnik 11:
85-94, illus.
(304) MOORE, P. J.
1961. THE RESPONSE OF SOILS TO DYNAMIC LOADINGS. Tech. Rpt. 7
by the Mass. Inst. Tech., to U.S. Army Corps of Engin.,
Waterways Expt. Sta., Vicksburg, Miss. 19 pp., illus.
(305) MORTON, C. T., and BUCHELE, W. F.
1960. EMERGENCE ENERGY OF PLANT SEEDLINGS. Agr. Engin 41-
428-431, 453-455, illus.
(306) MURAYAMA, S., and HATA, S.
1957. ON THE EFFECT OF REMOLDING CLAY. 4th Internatl. Conf. Soil
Mech. and Found. Engin. Proc. (London) 1: 80-82, illus.
(307) MURNAGHAN, F. D.
1951. FINITE DEFORMATION ON AN ELASTIC SOLID. 140 pp., illUS.,
New York.
(308) MURPHY, G.
1950. SIMILITUDE IN ENGINEERING. 302 pp., New York
(309) NAKAYA, U., TADA, M., SEKIDO, Y., and TAKANO, T.
1936. THE PHYSICS OF SKIING. Hokkaido ' Imper. Univ. Faculty
Sei. Jour. pp. 265-287, illus.
(310) NATIONAL INSTITUTE AGRICULTURAL ENGINEERING, SILSOE, ENGLAND.
1954. SINGLE WHEEL TESTER. Rpt. 40, 23 pp., illus.
(311) NATIONAL TILLAGE MACHINERY LABORATORY, AUBURN, ALA.
Unpublished Data.
(312) NICHOLS, M. L.
1925. THE SLIDING OF METAL OVER SOIL. Agr. Engin. 6: 80-84, illus.
(313)
1929. METHODS OF RESEARCH IN SOIL DYNAMICS; AS APPLIED TO
IMPLEMENT DESIGN. Ala. Agr. Expt. Sta. Bul. 229, 28 pp
illus.
(314)
1930. DYNAMIC PROPERTIES OF SOIL AFFECTING IMPLEMENT DESIGN
Agr. Engin. 11: 201-204, illus.
(315)
1931. THE DYNAMIC PROPERTIES OF SOIL. I., AN EXPLANATION OF
THE DYNAMIC PROPERTIES OF SOIL BY MEANS OF COLLOIDAL
FILMS. Agr. Engin. 12: 259-264, illus.
(316)
1931. THE DYNAMIC PROPERTIES OF SOIL. 11. SOIL AND METAL FRIC-
TION. Agr. Engin. 12: 321-324, illus.
(317)
1932. THE DYNAMIC PROPERTIES OF SOIL. III. SHEAR VALUES OF
UNCEMENTED SOILS. Agr. Engin. 13: 201-204, illus
(318) and BAVER, L. D.
1930. AN INTERPRETATION OF THE PHYSICAL PROPERTIES OF SOIL
AFFECTING TILLAGE AND IMPLEMENT DESIGN BY MEANS OF
ATTERBERG CONSISTENCY CONSTANTS. 2nd Internatl. Cong.
Soil Sei. Proc. (Leningrad-Moscow), pp. 175-188, illus
(319) and KUMMER, T. H.
1932. THE DYNAMIC PROPERTIES OF SOIL. IV. A METHOD OF ANALY-
SIS OF PLOW MOLDBOARD DESIGN BASED UPON DYNAMIC PROPER-
TIES OF SOIL. Agr. Engin. 13: 279-285, illus.
(320) and RANDOLPH, J. W.
1925. A METHOD OF STUDYING SOIL STRESSES. Agr. EugiU 6' 134-
135, illus.
SOIL DYNAMICS IN TILLAGE AND TRACTION 477
(321) and REAVES, C. A.
1955 SOIL STRUCTURE AND CONSISTENCY IN TILLAGE IMPLEMENT DE-
SIGN. Agr. Engin. 36: 517-520, 522, illus.
(322) and REAVES, C. A. .
1958. SOIL REACTION : TO SUBSOILING EQUIPMENT. Agr. Engin. 39 :
340-343, illus.
(323) and REED, I. F.
1934. SOIL DYNAMICS, VI. PHYSICAL REACTIONS OF SOILS TO MOLD-
BOARD SURFACES. Agr. Engin. 15: 187-190, illus.
(324) REED, I. F., and REAVES, C. A.
1958. SOIL REACTION : TO PLOW SHARE DESIGN. Agr. Engin. 39 :
330-339, illus.
(325) NIKIFOROFF, C. C. c M d .
1941. MORPHOLOGICAL CLASSIFICATION OF SOIL STRUCTURE. SOll SCl.
52: 193-211, illus.
(326) ODELL, R. T., THORNBURN, T. H., and MCKENZIE, L. J.
1960. RELATIONSHIPS OF ATTERBERG LIMITS TO SOME OTHER PROPER-
TIES OF ILLINOIS SOILS. Soil Sci. Soc. Amer. Proc. 24: 297-
300, illus.
(327) PATTY, R. L., and MINIUM, L. W.
1945. RAMMED EARTH WALLS FOR FARM BUILDINGS. S.D. Agr. Expt.
Sta. Bui. 277, 63 pp., illus.
(328) PAYNE, P. C. J.
1956. A FIELD METHOD OF MEASURING SOIL-METAL FRICTION. JOUr.
Soil Sei. 7: 235-241, illus.
(329)
1956. THE RELATIONSHIP BETWEEN THE MECHANICAL PROPERTIES OF
SOIL AND THE PERFORMANCE OF SIMPLE CULTIVATION IMPLE-
MENTS. Jour. Agr. Engin. Res. 1: 23-50, illus.
(330)
1956. WINCH SPRAG DESIGNED TO UTILIZE SOIL FRICTION. JOUr.
Agr. Engin. Res. 1: 51-55, illus.
(331) and FOUNTAINE, E. R.
1954. THE MECHANISM OF SCOURING FOR CULTIVATION IMPLEMENTS.
Nati. Inst. Agr. Engin. Tech. Memo. 116, 11 pp., illus.
(332) and TANNER, D. W.
1959. THE RELATIONSHIP BETWEEN RAKE ANGLE AND THE PERFORM-
ANCE OF SIMPLE CULTIVATION IMPLEMENTS. Jour. Agr. Eugiu.
Res. 4: 312-325, illus.
(333) PEATTIE, K. R., and SPARROW, R. W.
1954. THE FUNDAMENTAL ACTION OF EARTH PRESSURE CELLS. JOUr.
Mech. Of Phys. of Solids. 2: 141-155, illus.
(334) PFOST, H. B.
1948. ANALYSIS OF GEOMETRICAL DESIGN FACTORS INFLUENCING SOIL
MOVEMENT OVER PLOW MOLDBOARD SURFACES. M. S. ThesiS,
Auburn Univ., 49 pp., illus.
(335) PHILLIPS, J. R.
1961. THE POWERED VEHICULAR WHEEL PLANE-ROLLING IN EQUILIB-
RIUM : A CONSIDERATION OF SLIP AND ROLLING RESISTANCE.
1st Internatl. Conf. Mech. Soil-Vehicle Systems Proc. (Tu-
rin), pp. 540-557, illus.
(336) PiLLSBURY, R. D., JR.
1960. APPLICATIONS OF FLUOROCARBON RESINS IN FARM EQUIPMENT.
Agr. Engin. 41: 802-803, 825, 826, illus.
(337) PLANTEMA, G. A.
1953. SOIL PRESSURE CELL AND CALIBRATION EQUIPMENT. 3rd In-
ternatl. Conf. Soil Mech. and Found. Engin. Proc (Zurich)
1: 283-288, illus.
(338) POLLARD, W. S., JR.
1951. SOIL KNOWLEDGE CAN PROVE USEFUL TO DESIGNERS. S.A.E.
Jour. 59 (9) : 39^1, 50, illus.
(339) RADFORTH, N. W.
1961. LAND FACTORS AND VEHICLE DESIGN IN OPERATIONS ON ORGANIC
TERRAIN. 1st Internatl. Conf. Mech. Soil-Vehicle Systems
Proc. (Turin), pp. 158-170, illus.
478 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(340) RANDOLPH, J. W.
1926. TRACTOR LUG STUDIES ON SANDY SOIL. Agr. Engin. 7: 178-
184, illus.
(341) and REED, I. F.
1938. TESTS OF TILLAGE TOOLS. II. EFFECTS OF SEVERAL FACTORS ON
THE REACTIONS OF FOURTEEN-INCH MOLDBOARD PLOWS. Agr.
Engin. 19: 29-33, illus.
(342) RATH JE, J.
1932. THE CUTTING OF SAND. Engineering 134: 116-118, illus.
(343) REAVES, C. A.
1961. SIMILITUDE OF PLANE CECISELS IN ARTIFICIAL SOILS. Amer.
Soc. Agr. Engin. Trans. 9: 147-150, illus.
(344) and COOPER, A. W.
1960. STRESS DISTRIBUTION IN SOILS UNDER TRACTOR LOADS. Agr.
Engin. 41: 20-21, 31, illus.
(345) and NICHOLS, M. L.
1955. SURFACE SOIL REACTION TO PRESSURE. Agr. Engin. 36: 813-
816, 820, illus.
(346) REDSHAW, S. C.
1954. A SENSITIVE MINIATURE PRESSURE CELL. Jour. Sei. Inst.
31: 467-469, illus.
(347) REECE, A. R.
1961. A THREE POINT LINKAGE DYNAMOMETER. Jour. Agr. EugiU.
Res. 6: 45-50, illus.
(348) REED,I. F.
1940. RESULTS OF LEGUME COVERAGE STUDIES. Agr. Eugiu 21 * 129-
130, 134, illus.
(349)
1940. A METHOD OF STUDYING SOIL PACKING BY TRACTORS. Agr.
Engin. 21: 281-282, 285, illus.
(350)
1941. TESTS OF TILLAGE TOOLS. IIL EFFECT OF SHAPE ON THE DRAFT
OF 14-INCH MOLDBOARD PLOW BOTTOMS. Agr. Eugiu. 22: 101-
104, illus.
(351)
1944. SOME FACTORS AFFECTING DESIGN OF TILLAGE MACHINERY AND
PROPOSED APPROACH TO THEIR EVALUATION. Soil Sci. Soc.
Amer. Proc. 9: 233-235, illus.
(352)
1948. PLOW SECTION STUDIES. Nati. Tillage Mach. Lab. (Auburn
Ala.) Ann. Rpt.
(353)
1948. BASIC STUDIES OF DISK BLADES FOR AGRICULTURAL IMPLEMENTS.
Nati. Tillage Mach. Lab. (Auburn, Ala.) Ann. Rpt.
(354)
1958. TRACK STUDIES. Nati. Tillage Mach. Lab. (Auburn, Ala.)
Ann. Rpt.
(355)
1958. MEASUREMENT OF FORCES ON TRACK TYPE TRACTOR SHOES.
Amer. Soc. Agr. Engin. Trans. 1: 15-18, illus.
(356)
1962. THE EFFECTS OF INFLATION PRESSURE, LOAD, AND DRAWBAR
PULL ON AXLE HEIGHT AND ROLLING RADIUS OF SIX TIRES.
Amer. Soe. Agr. Engin. Trans 5: 125, 132, illus.
(357) and BERRY, M. O.
1946. TESTS COMPARING PLOW BOTTOMS. Nati. Tillage Mach. Lab.
Ann. Rpt, Auburn, Ala.
(358) COOPER, A. W., and REAVES, C. A.
1959 EFFECTS OF TWO-WHEEL AND TANDEM DRIVE TRACTION AND SOIL
COMPACTING STRESSES. Amer. Soc. Agr. Engin. Trans 2-
22-25, illus. * '
(359) and GORDON, E. D.
1951. DETERMINING THE RELATIVE WEAR RESISTANCE OF METALS.
Agr. Engin. 32: 9^-100, illus.
SOIL DYNAMICS IN TILLAGE AND TRACTION 479
(396)
1953. [PRESSURE DISTRIBUTION IN THE SOIL AND SOIL DEFORMATION
UNDER TRACTOR TIRES.] Grundlagen der Landtechnik 5: 49-
63, illus.
(397)
1953. [FRICTION AND COHESION IN ARABLE SOILS.] Grundlagen der
Landtechnik 5: 64-80, illus.
(398)
1956. [SOME PRINCIPLES OF SOIL MECHANICS AS APPLIED TO AGRI-
CULTURAL ENGINEERING.] Grundlagen der Landtechnik 7:
11-27, illus. (Nati. Inst. Agr. Engin. Eng. Translation 53.)
(399)
1957. [INFLUENCE OF THE SHAPE AND ARRANGEMENT OF THE TOOL
UPON THE DRIVING TORQUE IN ROTARY CULTIVATORS.] Grund-
lagen der Landtechnik 9: 69-87, illus.
(400)
1957. [IMPROVEMENT OF TRACTOR STEERING ON SIDE SLOPES BY
MEANS OF DISC COULTERS.] Grundlagen der Landtechnik 9:
113-118, illus.
(401)
1958. FUNDAMENTALS OF PRESSURE DISTRIBUTION AND SOIL COMPAC-
TION UNDER TRACTOR TIRES. Agr. Eugiu. 391 276-281, 290,
illus.
(402)
1959. [INVESTIGATIONS ON THE SHAPE OF PLOUGH BODIES FOR HIGH
SPEEDS.] Grundlagen der Landtechnik 11: 22-39, illus.
(Nati. Inst. Agr. Engin., Eng. Translation 87.)
(403)
1960. SUITING THE PLOW BODY SHAPE TO HIGHER SPEEDS. Grund-
lagen der Landtechnik 12: 51-62, illus. [Nati. Inst. Agr.
Engin, Eng. Translation 101.]
(404) CHANCELLOR, W. J., and SCHMIDT, R. H.
1959. SOIL DEFORMATION AND COMPACTION DURING PISTON SINKAGE.
Amer. Soc. Agr. Engin. Trans. 5: 235-239. 1962.]
(405) and EGGENMULLER, A.
1959. [FAST-RUNNING ROTARY CULTIVATORS AND SLOW-RUNNING DIG-
GERS : INVESTIGATIONS ON INDIVIDUAL TOOLS.] Grundlagen der
Landtechnik 11: 72-80, illus. Nati. Inst. Agr. Engin. Eng.
Translation 111.]
(406) and THIEL, R.
1957. [TECHNICAL PROBLEMS WITH ROTARY CULTIVATORS.] Grund-
lagen der Landtechnik 9: 39-49, illus.
(407) SOURISSEAU, J. H.
1935. [DETERMINATION AND STUDY OF PHYSICO-MECHANICAL PROPER-
TIES OF SOIL.] Organ, and Raps, du II Cong. Internatl. de
Genie Rural (Madrid), pp. 159-194, illus.
(408) SOUTHWELL, P. H.
1964. AN INVESTIGATION OF TRACTION AND TRACTION AIDS. Amer.
Soc. Agr. Engin. Trans. 7: 190-193, illus.
(409) STANILAND, L. N.
1959. FLUORESCENT TRACER TECHNIQUES FOR THE STUDY OF SPRAY
AND DUST DEPOSITS. Jour. Agr. Engin. Res. 4: 110-125,
illus.
(410) STAUFFER, L. H.
1927. MEASUREMENT OF PHYSICAL CHARACTERISTICS OF SOILS. SOIL
Sei. 24: 373-379, illus.
(411) STEINBRUEGGE, G.- W.
1961. THE IDEAL TRACTIVE EFFICIENCYOF SOILS. 1st Internatl. Conf.
Mech. Soil-Vehicle Systems Proc. (Turin) pp. 558-566, illus.
(412) STEWART, A.
1948. A TUMBLING TEST FOR PRESSED PELLETS. Jour. Sci. lUSt. 25 :
438-440, illus.
482 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(433)
1960. SCALED VEHICLE MOBILITY FACTORS. Second Interim Rpt.
Proj. 9R97-40-001-01, House Task 4.2, Ö4 pp., illus.
(434)
1960. VTOL DOWNWASH IMPINGEMENT STUDY SURFACE EROSION TESTS.
Transport. Res. Com. Tech. Rpt. 60-67, 174 pp., illus.
(435)
1961. SCALE MODEL OF VEHICLES IN SOILS AND SNOWS. ReS. Tech.
Memo. 33, 33 pp., illus.
(436) TROUSE, A. C, and HUMBERT, R. P.
1961. SOME EFFECTS OF SOIL COMPACTION ON THE DEVELOPMENT OF
SUGAR CANE ROOTS. Soil Sci. 91: 208-217, illus.
(436) TRUITT, T. D., and ROGERS, A. E.
1960. BASICS OF ANALOG COMPUTERS, illus., New York.
(438) TRULLINGER, R. W.
1924. RESEARCH IN AGRICULTURAL ENGINEERING 1923. Agr. Engin.
5: 107-111.
(439)
1924. EVOLUTION AND PROGRESS OF AGRICULTURAL ENGINEERING AT
THE AGRICULTURAL EXPERIMENT STATIONS. Agr. EugiU. 5 !
276-279.
(440)
1925. SOIL COLLOIDS AND TILLAGE. Agr. Eugiu. 6: 61-63, 84-87.
(441)
1926. RESEARCH IN AGRICULTURAL ENGINEERING—1925. Agr. Engin.
7: 279-282.
(442) TURNBULL, W. J., JOHNSON, S. J., and MAXWELL, A. A.
1949. FACTORS INFLUENCING COMPACTION OF SOILS. Nat'l. Res.
Council Highway Res. Board Bui. 23, illus.
(443) TYURIN, I. V.
1957. THE MAL'TSEV SYSTEM OF CULTIVATION. Soils and Fert. 20:
305-309.
(444) UFFELMANN, F. L.
1956. VEHICLE MOBILITY RESEARCH IN THE UNITED KINGDOM. Land
Locomotion Lab. (Detroit) Res. Rpt. 3, pp. 5-19, illus.
(445) UMEDA, S.
1958. CHARACTERISTICS OF SOIL RESISTANCE OF ROTARY TILLAGE TINES.
Univ. Osaka [Japan] Bui., Ser. B, v. 8, pp. 9-17, illus.
(446) UNITED STATES DEPARTMENT OF AGRICULTURE.
1938. SOILS AND MEN. U.S. Dept. Agr. Yearbook 1938. 1232 pp.,
illus.
(447)
1951. SOIL ^SURVEY MANUAL. U.S. Dept. Agr., Agr. Handb. 18,
503 pp., illus.
(448)
1954. DIAGNOSIS AND IMPROVEMENT OF SALINE AND ALKALI SOILS.
U.S. Dept. Agr., Agr. Handb. 60, 160 pp., illus.
(449) , Soil Survey Staff.
1960. SOIL CLASSIFICATION SYSTEM. 265 pp., illUS.
(450) UNITED STATES DEPARTMENT OF THE INTERIOR, BUREAU OF RECLAMATION.
1960. EARTH MANUAL. 751 pp., illus. Denver.
(451) VAIGNEUR, H. O.
1959. LOADING CHARACTERISTICS OF PLASTIC LINED MOLE DRAINS.
M. S. Thesis, Clemson College, illus,
(452) VAN BAVEL, C. H. M.
1949. MEAN WEIGHT-DIAMETER OF SOIL AGGREGATES AS A STATISTICAL
INDEX OF AGGREGATION. Soil Sci. Soc. Amer. Proc. 14: 20-
23, illus.
(453)
1960. SOIL DENSITY MEASUREMENT WITH GAMMA RADIATION. 7th
Internatl. Cong. Soil Sci. Trans. 1: 284-289, (Madison, Wis.)
illus.
(454) UNDERWOOD, N., and RAGER, S. R.
1957. TRANSMISSION OF GAMMA RADIATION BY SOILS AND SOIL DEN-
siTOMETRY. Soil Sci. Soc. Amcr. Proc. 21: 588-591, illus.
484 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(455) VANDEN BERG, G. E,
1961. ANALYSIS OF FORCES ON TILLAGE TOOLS. U.S. Nati. Tillage
Mach. Lab. (Auburn, Ala.) Ann. Rpt.
(456)
1962. CONTINUOUS ANALOG TECHNIQUES IN EXPERIMENTAL RESEARCH.
Paper presented South. A.gr. Workers Meeting, Jacksonville,
Fla., 7 pp., illus.
(457)
1962. TRIAXIAL MEASUREMENTS OF SHEARING STRAIN AND COMPAC-
TION IN UNSATURATED SOIL. Auier. Soc. Agr. Engin., Paper
62-648.
(458)
1962. REQUIREMENTS FOR A SOIL MECHANICS. Amer. Soc. Agr.
Engin. Trans. 4: 234-238.
(459) BucHELE, W. F., and MALVERN, L. E.
1958. APPLICATION OF CONTINUUM MECHANICS TO SOIL COMPACTION.
Amer. Soc. Agr. Engin. Trans. 1: 24-27, illus.
(460) and GILL, W. R.
1962. PRESSURE DISTRIBUTION BETWEEN A SMOOTH TIRE AND THE
SOIL. Amer. Soc. Agr. Engin. Trans. 5: 105-107, illus.
(461) and REED, I. F.
1962. TRACTIVE PERFORMANCE OF RADIAL PLY AND CONVENTIONAL
TRACTOR TIRES. Amer. Soc. Agr. Engin. Trans. 5: 126-129,
132, illus.
(462) REED, I. F., and COOPER, A. W.
1961. EVALUATING AND IMPROVING PERFORMANCE OF TRACTION DE-
VICES. 1st Internatl. Conf. Mech. Soil-Vehicle System Proc,
(Turin), pp. 402-411, illus.
(463) VASEY, G. H., and NAYLOR, I. T.
1958. FIELD TESTS ON 14-30 TRACTOR TYRES. Jour. Agr. Engin. Res
3: 1-8, illus.
(464) and NAYLOR, I. T.
1958. FIELD TESTS ON 14-30 TRACTOR TYRES. II. STATISTICAL
TREATMENT OF RESULTS. Jour. Agr. Eugin. Res. 3: 187-198,
illus.
(465) VETROV, YU. A.
1958. [FRICTION BETWEEN THE BLADE AND SOIL IN THE PROCESS OF
CUTTING.] Nauch. Dokl. Vysshei Shkoly [Stroilestro] 2:
111-124, illus. [Eng. Translation J. Crearer Lib. RTS 1445]
(466) VINCENT, E. T.
1961. PRESSURE DISTRIBUTION ON AND FLOW OF SAND PAST A RIGID
WHEEL. 1st Internatl. Conf. Mech. Soil-Vehicle Systems
Proc, (Turin), pp. 858-878, illus.
(467) VoMOciL, J. A., FouNTAiNE, E. R., and REGINATO, R. J.
1958. THE INFLUENCE OF SPEED AND DRAWBAR LOAD ON THE COM-
PACTING EFFECT OF WHEELED TRACTORS. Soil Sci. Soc. Amer
Proc. 22: 178-180, illus.
(468) and WALDRON, L. J.
1962. EFFECT OF MOISTURE CONTENT ON TENSILE STRENGTH OF UN-
SATURATED GLASS BEAD SYSTEMS. Soil Sci. Soc. Amer Proc
26: 409-412, illus.
(469) WALDRON, L. J., and CHANCELLOR, W. J.
1961. SOIL TENSILE STRENGTH BY CENTRIFUGATION. Soil Sci Soc
Amer. Proc. 25: 176-180, illus.
(470) WANN, R. L., and REED, I. F.
1962. STUDIES OF TRACTOR TIRE TREAD MOVEMENT. Amer Soc Agr
Engin. Trans. 5: 130-132, illus.
(471) WARLAM, A. A.
1946. STRESS-STRAIN AND STRENGTH PROPERTIES OF SOILS. D. SC
Thesis, Harvard Univ., illus.
(472) WATERWAYS EXPERIMENT STATION, VICKSBURG, MISS.
1948. DEVELOPMENT OF TESTING INSTRUMENTS. Tech. Memo 3-240
Trafficability of Soils, 3rd Sup. 66 pp., illus.
SOIL DYNAMICS IN TILLAGE AND TRACTION 485
(473)
1950. MISCELLANEOUS LABORATORY TESTS. Tech. Memo. 3-271, Soil
Compaction Investigation Report 5, 37 pp., illus.
(474)
1950. TESTS ON TOWED VEHICLES. Tech. Memo. 3-240, 7th Sup.
109 pp., illus.
(475)
1951. TRAFFiCABiLiTY OF SOILS, SLOPE STUDIES. Tech. Memo. 3-
240, 29 pp., illus.
(476)
1952. TORSION SHEAR APPARATUS AND TESTING PROCEDURES. Bul.
38, 76 pp., illus.
(477)
1953. UNIFIED SOIL CLASSIFICATION SYSTEM. Tech. Meuio. 3-357,
V. 1, 30 pp ; V. 2, 11 pp. ; V. 3, 9 pp.
(478)
1953. INVESTIGATIONS OF PRESSURES AND DEFLECTIONS FOR FLEXIBLE
PAVEMENTS. Tech. Memo. 3-323, Rpt. 3, 13 pp., illus.
(479)
1954. EFFECT OF SIZE OF FEET ON SHEEPSFOOT ROLLER. Tech. MemO.
3-271, Rpt. 6, 29 pp., illus.
(480)
1954. INVESTIGATIONS OF PRESSURES AND DEFLECTIONS FOR FLEXIBLE
PAVEMENTS. Tech. Memo. 3-323, Rpt. 4, Homogeneous Sand
Test Section, 60 pp., illus.
(481)
1955. PRESSURE CELLS FOR FIELD USE. Bul. 40, 33 pp., iUuS.
(482)
1955. EFFECT ON SOIL COMPACTION OF TIRE PRESSURE AND NUMBER OF
COVERAGES OF RUBBER-TIRED ROLLERS AND FOOT-CONTACT PRES-
SURE OF SHEEPSFOOT ROLLERS. Tech. Mcmo. 3-271, Rpt. 7,
41 pp., illus.
(483)
1958. STUDIES OF AERIAL CONE PENETROMETER. Tecll. Rpt. 3-462,
Rpt. 2, 19 pp., illus.
(484)
1959. DEFLECTION OF MOVING TIRES. Tech. Rpt. 3-516, Rpt. 1, 19
pp., illus.
(485)
1959. INVESTIGATION OF EFFECTS OF 50,000-LB. WHEEL-LOAD TRAFFIC
ON A SHALLOW-BURIED FLEXIBLE PIPE. Misc. Paper 4-364,
14 pp., illus.
(486)
1959. A LIMITED STUDY OF SNAP-TRACKS. Misc. Paper 4-322, 8 pp.,
illus.
(487)
1959. PILOT STUDY TO EVALUATE THE SQUEEZE TEST FOR USE IN
VEHICLE MOBILITY RESEARCH. Misc. Paper 4-350, 15 pp.,
illus.
(488)
1960. LIST OF PUBLICATIONS. 141 pp.
(489)
1960. STRESSES UNDER MOVING VEHICLES. Tech. Rpt. 3-545, Rpt. 3,
33 pp., illus.
(490)
1960. TRAFFICABILITY OF SNOW. Tech. Memo. 3-414, Rpt. 3, 65 pp.,
illus.
(491)
1960. PETROGRAPHIC DATA ON BEACH, DUNE AND RIVER SANDS. MiSC.
Paper 6-^08.
(492)
1961. SOIL TRAFFICABILITY CLASSIFICATION SCHEME. Ist lutematl.
Conf. Mech. Soil-Vehicle Systems Proc. (Turin), pp. 567-
574, illus.
486 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE
(493)
1961. TESTS WITH RIGID WHEELS. Tech. Rpt. 3-565, Rpt. 1, 35 pp.,
illus.
(494)
1961. TRAFFiCABiLiTY OF SOILS. Tech. Memo. 3-240, 16th Sup. 65
pp., illus.
(495)
1962. ABSTRACTS OF REPORTS PUBLISHED BY ARMY MOBILITY RE-
SEARCH CENTER TRAFFICABILITY AND MOBILITY RESEARCH.
22 pp.
(496)
1965.
WHEELS ON SOFT SOILS, AN ANALYSIS OF EXISTING DATA
Tech. Kept. 3-670, 81 pp., illus.
(497) WEAVER, H. A.
1950. TRACTOR USE EFFECTS ON VOLUME WEIGHT OF DAVIDSON LOAM
Agr. Engin. 31: 182—183, illus.
(498) and JAMISON, V. C.
1951. EFFECTS OF MOISTURE ON TRACTOR TIRE COMPACTION OF SOIL
Soil Sei. 71: 15-23, illus.
(499) WEBER, F.
1932.
[INVESTIGATIONS ON THE INFLUENCE OF ELECTRIC CURRENT ON
THE TRACTIVE FORCE REQUIRED IN PLOWING.] Thesis, Tech-
nischen Hochscule, München, 47 pp., illus
(500) WEIDENBAUM, S. S.
1918.
A STUDY OF THE PLOW BOTTOM AND ITS ACTION UPON THE
BURROW SLICE. Ph. D. Thesis, Cornell Univ., iilus
(504) WiENEKE, F.
1955.
THE MATHEMATICAL DETERMINATION OF THE RUNNING CHAR-
ACTERISTICS OF A POWER-DRIVEN TRAILER. Laudtechnische
Forsch. (Munich) 5: 26-29, illus. [Jour. Agr. Engin. Res
1: 101-105. 1956, Eng. Translation.]
(505) WiLLETTS, B. F.
1954. THE PERFORMANCE OF FOOTINGS ON, AND CULTIVATION IM-
PLEMENTS IN, SOILS. Ph. D. Thesis, Univ. of Durham
England, illus.
(506) WILSON, R. W.
1960. DISCUSSION OF DESIGN FACTORS IN A SPRING TRIP BEAM AS-
SEMBLY FOR MOLDBOARD PLOWS. Amer. Soc. Agr. Engin
Trans. 3(2): 64-65, illus. ^ ^
(507) WINKELBLECH, C. S.
1961. SOIL AGGREGATE SEPARATION CHARACTERISTICS OF SECONDARY
/^no^ ^r TILLAGE TOOLS. M. S. Thesis, Ohio State Univ., illus.
(508) WINTERKORN, H. F.
1936. SURFACE CHEMICAL FACTORS INFLUENCING THE ENGINEERING
^n^^r^Ti. T ^''íí: ^^^'^- ^^'- ^^^^^il' Highway Res.
S.T 1^^*^ ^^^' ^^^^ P^^^C" (Washington, D.C.) pp. 293-
301, illus.
(509) WOODRUFF, C. M.
1936. LINEAR CHANGES IN THE SHELBY LOAM PROFILE AS A FUNCTION
/^-.^x ..r ^^ ®^^^ MOISTURE. Soil Sei. Soc. Amer. Proc. 1: 65-70 illus
(510) WOODRUFF, N. P., and CHEPIL, W. S. ^. -L. OO íU, iiius.
1956. IMPLEMENTS FOR WIND EROSION CONTROL. Agr Euffin 37«
751-754, illus. & . oi .
(511) CHEPIL, W. S ., and LYNCH, R. D
1957. EMERGENCY CHISELING TO CONTROL WIND EROSION Kans
Agr. Expt. Sta. Tech. Bui. 90, 24 pp., illus.
SOIL DYNAMICS IN TILLAGE AND TRACTION 487
Abrasion, by airborne soil, 106, 107 ; described, 52 ; factors affecting, 110, 112 ;
mechanism, 109; of plastic plow covers, 235; parameters, 52; relation with
modulus of rupture, 106 ; standard test for soil. 111 ; see also wear
Acceleration, contribution to draft force, 318 ; in plow mechanics, 176 ; of soil
by blade, 317 ; of soil in soil-tool mechanics, 130 ; see also speed
Acceleration, soil, by moldboard plows, 228
Acceleration, tool, by increased forces, 213
Action, described by complete mechanics, 123; described by mechanics, 120;
determined by interest, 120; influenced by soil-tool geometry, 191; param-
eters deñned by mechanics, 218
Active behavior, basis for use, 300 ; caused by dominant forces, 55 ; empirical,
102; measured, 334
Adhesion, apparent, 40; defined by equation, 90; description, 42; effect of
angle of wetting, 48 ; effect of moisture tension, 45 ; effect of surface tension
on, 44; effect of temperature on, 48; effect of wetting plow, 236; effect of
sliding force, 162 ; effect on machine performance, 91 ; effect on tool filling,
252; equivalence of stress, 90, 162; equivalence of weight, 50; factors gov-
erning, 91; in soil-tool mechanics, 132, 141; interaction with friction, 49;
measurement of, 44; Nichols' phases of friction, 51; on bulldozer blades,
252 ; on moldboard plows, 48, 176 ; on tires, 412 ; on tool edges, 244 ; related
to physical properties, 89 ; scouring of plows, 91 ; soil body formation, 192 ;
stickiness, 89
Aeration, required for plant growth, 303
Aggregates, behavior under stress, 305
Aggregate size, and cutting force, 209; see also clods
Aggregate stability, method to assess, 105
Agricultural operation, cause compactness, 452
Agricultural system, hierarchy, 450; influenced by soil dynamics, 447
Air blast, to manipulate soil, 237
Air displacement, measure of volume strain, 30
Air plow, design, 236
All wheel drive, vehicles, 418
Analogs, applied to soil dynamics research, 458
Anchor, plant material in soil, 327 ; sprag in soil, 408
Angle, for minimum penetration resistance, 184; of soil body, 143; of shear
surface, 131; of tool orientation, 256; penetrometer and relations, 187
Angle, ascending, defined, 170
Angle of bulldozing blades, effect on draft, 251
Angle of cutter, effect on soil deformation, 154
Angle of internal friction, determination of, 67-68
Angle of sharpness, effect on plow draft, 248 ; for tiller tines, 275
Angle of sliding friction, of soils, 51
490
INDEX 491
Angle of tool, effect on scouring, 178; soil confinement, 200
Angle of wetting, effect of surface roughness, 49; measurement, 48; soil
solutions on metals, 49
Apparent coefficient of sliding friction, measurement, 49
Arch action, around stress transducers, 28; in compression chambers, 81; in
granular materials, 24; in soil channels, 239; in tool action, 203; inñuence
on traction performance, 405 ; observed in glass-sided box, 114
Artificial soils, to stabilize strength, 22
Area of contact, effect on adhesion, 47, 49 ; effect on penetrating force, 183 ;
effect on plow performance, 237 ; effect on sticking, 93 ; measured for tires,
364 ; of cutters, 152 ; of tires, 365 ; tire footprint, 364
Area of influence, of grousers, 405
Auxilliary traction devices, performance, 410
Edgeshape, design, 249 ; described, 242 ; effect of wear, 245 ; effect on plowing
depth, 247 ; effect on tool force, 243 ; effect on soil movement, 244 ; of tillage
tools, 221
Effective stress, during impacts, 439; in shear tests, 73-74
Efficiency, different plows, 338 ; tillage tool cuts, 272 ; tractive devices, 374
Elastic deformation, during cutting, 158
Elastic theory, restricted to small strain, 20-21
Elasticity, coefficient, 157
Elasticity equations, to calculate stress distributions, 26
Electrical power, see power
Electro-osmosis, to reduce adhesion, 269; requisite conditions, 268
Empirical behavior equations, development role, 217
Empirical mechanics, need, 63
Energy, basis of tool performance, 263; conserved by tool design, 325; effect
of worn shares, 247 ; impacts on clod size, 321 ; in U.S. tillage, 220 ; in U.S.
traction, 340 ; multiple sources for tools, 265 ; of blast streams, 109 ; possible
savings in tillage, 452 ; required for actions, 298 ; required for traction, 374 ;
to breakup soil, 263, 338 ; to shatter clods, 104
Environmental forces, effect on soil compactness, 431 ; effect on soil conditions,
302
Evaluation, based on desired performance, 336 ; basis for vehicle performance,
381; feedback for design, 299; of tillage for plant growth, 303; of tire
performance, 400 ; of traction performance, 365, 379 ; of vehicle performance,
381
Evaluation factors, described, 337
Evaluation of performance, described, 334-339; principles, 299
Evaluation of tillage,* plow performance, 338
Evaluation of traction, comparative techniques, 385
Parameters, abrasion, 52 ; for soil-tool mechanics, 123 ; see also cohesion, fric-
tion, tool, tire
Partial mechanics, development, 125; utility, 160
Particle size, coefficient of friction, 112 ; effect on abrasion, 106 ; effect on blast
penetration, 108-109 ; effect on mixing, 323 ; effect on sliding path, 171 ; effect
on wind erodability, 106
Passive behavior, basis for soil use, 300 ; transmitted force systems, 55
Passive earth pressure theory, in sprag equation, 408; in tillage tool me-
chanics, 139
Path of motion, drawn tools, 260 ; oscillating tools, 283 ; rotary tools, 275
Path of soil sliding, apparatus for tracing, 229 ; method to describe, 229
Path of tool travel, description, 269
Penetration, composite behavior, 181
Penetration force, calculated, 184
Penetration resistance, a composite property, 94; influenced by geometry, 182
Penetrometer(s), aerial, 95; impact parameter c, 94; recording, 95; static
parameter, 94 ; tilting plate, 100 ; to assess soil conditions, 189 ; to character-
ize soil, 94 ; to predict plow draft, 96 ; to predict rolling resistance, 96, 424 ;
to predict seedling emergence, 96; types, 94
Performance, based on clod size, 103 ; based on soil conditions, 336 ; calculated
with mechanics, 298; criterion for tractive devices, 366; defined, 298; de-
scribed, 298; measures, 377; measuring soil condition, 310; of material
placement machines, 330; of moldboard plow, 336; of rotating plow, 336; of
rotary tillers, 271; of spading machine, 336; of tillage tools for plant
growth, 303; of traction devices, 365; of winch sprags, 409; qualitative
descriptions, 299; specific goals, 300
Performance criterion, soil factors, 298
Performance factors, described, 337
Permanent deformation, as a yield criterion, 32 ; of soil, 21
Philosophy, for tillage, 454
Photo elasticity, use in soil, 29
Physical properties, changes described by mechanics, 304; circumvented by
behavior equations, 63; effect on abrasion, 106; related to behavior prop-
erties, 59; relation with behavior, 30^308
Pi-terms, in dimensionless equations, 426
Pipe, installed by tillage, 331
Placement of plant residue, by plowing techniques, 328
500 AGRICULTURE HANDBOOK 316, U.S. DEPT. OF AGRICULTURE