You are on page 1of 171

TRACTOR-I M PLEM ENT

SYSTEMS
TRACTOR-IMPLEMENT
SYSTEMS

Ralph Alcock

Agricultural Engineering Department


South Dakota State University
Brookings, South Dakota


aVI AVI PUBLISHING CO., INC.
Westport, Connecticut
© Copyright 1986 by
Softcover reprint of the hardcover 1st edition 1986
THE AVI PUBLISHING COMPANY, INC.
250 Post Road East
P.O. Box 831
Westport, Connecticut 06881

All rights reserved. No part of this work


covered by the copyright hereon may be
reproduced or used in any form or by any
means-graphic, electronic, or mechanical,
including photocopying, recording, taping, or
information storage and retrieval systems-
without written permission of the publisher.

Library of Congress Cataloging-in-Publication Data

Alcock, Ralph.
Tractor-implement systems.

Includes bibliographies and index.


1. Farm tractors. 2. Agricultural implements.
I. Title.
S711.A44 1986 631.3'72 85-28629
ISBN-13: 978-1-4684-6881-6 e-ISBN-13: 978-1-4684-6879-3
001: 10.1007/978-1-4684-6879-3

ABCDE 5432109876
Contents

Preface ix
Acknowledgments xi

1 Fundamental Considerations

1.1 Introduction
1.2 Systems 5
1.3 Energy and Principles of Dynamics 6
1.4 The Conservation of Energy 8
References 9

2 Thermodynamics

2.1 The Steady-Flow Energy Equation 10


2.2 The N onflow Energy Equation 13
2.3 The First Law of Thermodynamics 14
2.4 The Second Law of Thermodynamics 14
v
vi Contents

2.5 The Thermodynamic Engine 15


2.6 Gas Laws 16
2.7 Specific Heat 18
2.8 Gas Processes 19
2.9 The Carnot Cycle 23
References 23

3 Power Measurement

3.1 Power 24
3.2 Torque 25
3.3 Indicated Power (IP) 27
3.4 Energy Balance 30
3.5 Measurement of Fuel Flow 33
3.6 Measurement of Air Flow 34
3.7 Tractor Performance 35
3.8 Drawbar Pull (F D ) 36
3.9 Nebraska Tractor Tests 41
References 42

4 Traction

4.1 Traction in Soils 44


4.2 Shear Strength 44
4.3 Plastic Limit 46
4.4 Coulomb and Micklethwaite Equations 46
4.5 Bekker Theory 49
4.6 Dimensional Analysis 53
4.7 Tractive Efficiency (11D) 54
4.8 Mobility Number 56
References 58

5 Tires and Ballasting for Agricultural Tractors

5.1 Tire Development 59


5.2 Tire Function 59
Contents vii

5.3 Tire Terminology 60


5.4 Ballasting 61
5.5 Tire Selection 68
References 72

6 Machinery Management

6.1 Fixed Costs 73


6.2 Operating Costs 76
6.3 Reliability 78
6.4 Selection of Machinery Capacity 83
6.5 Replacement Theory 86
References 91

7 Vehicle Performance

7.1 Drawbar Performance 92


7.2 Engine Torque Requirement 97
7.3 Vehicle Speed 97
7.4 Vehicle Acceleration 98
7.5 Longitudinal Stability 99
7.6 Tractor Dynamics 101
7.7 Tractor-Implement Forces 103
7.8 Tractor-Implement Dynamics 104
References 108

8 Connecting Tractor and Implement (Sverker


Persson)

8.1 Pull-Type, Semimounted,


or Mounted Implement 110
8.2 Hitch Point 111
8.3 Load Transfer Due to Pull 113
8.4 Effect of Load Transfer 116
8.5 Free-Link Three-Point Hitches 120
8.6 Powered Three-Point Hitch 121
8.7 Motions of Fully Mounted Implement
in Three-Point Hitch 122
viii Contents

8.8 Lifting Capacity of the Hydraulic System 123


8.9 Draft-Sensing Hydraulic Systems 123
8.10 Position Control 124
8.11 Hydraulic Systems 125
8.12 Open- or Closed-Center Hydraulics 125

9 Operating the Tractor and the Implement


(Sverker Persson)

9.1 The Tractor as a Workplace 127


9.2 Tractor Controls and Instruments 130
9.3 Operator Environment 134
9.4 Operator Safety 136
9.5 Systems for Hitching the Implement
to the Tractor 137
References 140

10 Developments in the "Tractor" Concepts

10.1 Introduction 141


10.2 Controlled-Traffic Farming 143
10.3 Battery-Powered Tractor 146
10.4 Systems Tractor 152
References 153

Appendix: Problems and Problem-Solving 154

Index 159
Preface

Agriculture has benefited considerably from the wide-scale use of


tractors and associated implements. Tractors have developed along
two, at times contradictory, design paths. The first of these has resulted
in functional improvements to the vehicle, making it capable of im-
proved performance in the often harsh environmental conditions in
which it is required to operate. For example, agricultural tires have
improved the versatility of tractors by allowing them to operate at
relatively high speeds on a variety of terrains, and hydraulics have
provided for flexibility in controlling and operating implements. It can
also be seen that these particular functional design improvements have
contributed significantly to the second design pathway, that of ergon-
omics, and the human-machine interface. Recent stress on the working
environment for the tractor operator has led to design improvements
relative to tractor cabs, the placement and labeling of controls, etc.
This text discusses those factors relevant to the design, selection,
and operation of tractor-implement systems. The audience for which
it is intended is undergraduate and graduate students of agricultural
engineering studying power and machinery. However, the text is suf-
ficiently applied to have relevance for those extension personnel in-
volved with advising farmers on the selection and operation of tractor-
implement combinations.
The tractor cannot be regarded in isolation from the implement, nor
from the environment in which both the tractor and implement are
ix
x Preface

working. The tractor-implement combination forms a system consist-


ing of a set of interacting components, with environmental constraints
provided by the terrain and atmosphere. The objective of this system
can be simply defined as the performing of a specific task to some
measurable standard. However, this approach fails to recognize that
the tractor-implement combination is itself a subsystem of the farming
enterprise. Thus, while it is important to analyze the subsystem in terms
of the power unit and traction fundamentals, we must also recognize
the importance of integrating the performance of this subsystem into
the goals of the major enterprise. For these reasons, this text develops
on the one hand from essentially isolated topics on thermodynamics,
and power measurement, and goes on to subsystem integration, for
example, tractor-implement matching, optimum capacity selection,
and replacement theory. On the other hand, later chapters focus on
interesting developments relating to ergonomics, electric tractors and
hitching systems.
The text is unique in the area of power and machinery in that it
develops the concept of a systems approach to tractor-implement com-
binations, while still retaining a useful content relevant to the funda-
mentals of tractor design and operation. In the appendix, relatively
unstructured problems that can be solved by computer are presented.
These are developed so as to require a progressively more detailed
level of programming.
Acknowledg ments

The author's appreciation and thanks are expressed to Dr. Mylo


Hellickson, Head of the Agricultural Engineering Department at South
Dakota State University, for his encouragement and support on this
project.
My thanks are also extended to Ingrid Werner for typing the man-
uscript, and to Kyle Elenkiwich for drafting the illustrations.

xi
1
Fundamental
Considerations

1.1 INTRODUCTION

A tractor is intended to provide transport functions for movement


of materials and for field work capabilities to power and propel agri-
cultural machines. In transport mode, the tractor is required to operate
with relatively low drafts over a wide range of operating speeds. For
field work, the tractor is often used for tasks requiring high drawbar
forces at speeds of 4-12 km hr-l. Thus, the tractor must interface with
the implement in providing the operating parameters necessary to meet
the performance objectives for both transport and field work functions.
The interface between tractor tire and the working surface consti-
tutes a functional limitation to the operation of a tractor-implement
system, as does the interface between the implement and the crop or
the soil. Other functional limitations are inherent in the design of the
tractor. For example, the transmission system will dictate the range
of operating speeds that can be achieved, and the noise levels or ride
comfort provided for the operator may limit the performance of the
tractor.
An example of a tractor-implement selection process is shown as
Fig. 1.1. If we neglect the design limitations to the functional perform-
2 1. Fundamental Considerations

Farm
system --------1---
I
I
+
Crop
profitability
I
I
I
I
I

Timeliness

Labor costs

Task costs

Implement availability

Tractor availability

Tractor-implement
costs ------+------

Minimal system
cost

Fig. 1.1. Tractor-implement selection iteration process.


1.1 Introduction 3

ance of the tractor, then tractor size or type is determined as that which
will minimize system cost while meeting the field capacity require-
ments. The figure shows the important decision steps to be made and
incorporates the factors likely to influence these processes. Despite its
versatility, the tractor and associated implement must compete with
the increasing availability of specialized self-propelled equipment. The
implement size required will be dictated not only by the field capacity
but also by the availability ofthe implement based on competing crops,
activities, and downtime. The same reasoning can be applied to the
tractor; however, it is important to recognize that the tractor may be
required to supply hydraulic power to the selected equipment.
In order to minimize system costs, or in order to compare custom
operation and leasing costs, it may be necessary to iterate the selection
process, using modified input criteria. The tractor is a complex and
expensive item of farm equipment, and as such it must provide the
performance required at a minimum level of cost to the system. Farm-
ers have little control over the price of tractors and implements; they
can, however, resort to logical selection methods, and adopt operating
and management procedures that will maximize return on investment.
Tractor designers have two basic options: (1) They can design for a
specialized market (for example, the row crop tractor; the small garden-
sized vehicle; or the large articulated four-wheel drive field vehicle).
(2) They can design tractors that provide for versatility in that they
can be used for a wide spectrum of tasks ranging from low-speed field
work to road transport. A unique example of the second option is the
Trantor® vehicle, which is manufactured in England. This dual-purpose
tractor is capable of traveling at a road speed of 80 km hr- I and has
a minimum field speed of approximately 1.5 km hr- I. The cab is sus-
pension mounted and capable of carrying three people. It is fitted with
a three-point linkage system, a two-speed' 'live" power take off (PTO) ,
and a pneumatically sprung drawbar. An earlier attempt at versatility
was the fitting of two tractors in a tandem arrangement so as to provide
for a greater power range and yet retain the flexibility provided by the
smaller individual units.
The Trantor® vehicle concept is somewhat against the recent trends
of increased tractor size, both in terms of mass and power. Yahya and
Goering (1977) analyzed the available Nebraska tractor test reports
over the period 1920-1975 and identified some key parameters with
respect to tractor design trends. Their results showed that, although
tractors became smaller and lighter in the 1920s and 1930s, since the
4 1. Fundamental Considerations

end of World War II tractor mass and power have risen rapidly. They
also showed that the power-to-mass ratio has been rising steadily in
this same period. In 1920, a typical tractor mass was 5600 kg, with a
drawbar power of 28 kW. The power-to-mass ratio was approximately
5.9 W kg - I . By 1940, the typical tractor mass had dropped to 3100 kg,
and a typical drawbar power was 22 kW. The power-to-mass ratio had
risen to approximately 6.5 W kg-I. In the period 1950-1970, tractor
mass increased rapidly. By 1970, the typical tractor mass was 6500 kg,
and the typical drawbar power had risen appreciably to a value of 58
kW. The power-to-mass ratio was now approximately 9 W kg-I.
Buchele (1959) suggested that a manufacturer with only two models
of tractor could effectively meet the majority of farm power require-
ments by spanning the 22-67 kW power range. At the opposite end of
the spectrum, specialization may result in power units that bear little
or no resemblance to the tractors currently used. Powered gantries
capable of spanning 15-20 m of crop have been proposed by a number
of researchers. For example, Matthews (1982) envisaged such a system
with a capability of supplying an intricate array of information to the
operator, such as depth, seed spacing and population, and soil mois-
ture. Chamen et ai. (1980) proposed that the gantry would be equipped
with a power unit of 100-150 kW, and be fitted with associated hy-
draulics and electrics. Implements would then be matched to the gantry
in modular form.
The applicability and feasibility of such systems is likely to be
strongly influenced by the efficiency with which they use fossil fuel
energy, or perhaps by their efficiency and flexibility in being adaptable
to electrical energy. Appropriate technology is defined by that overlap
between what is technically feasible and what is socially desirable.
The agricultural industries ofthe world's developed nations consume
a small portion of their commercial energy usage. A typical estimate
for the amount of energy used by the agricultural industry in the United
States is 2.8%, with figures ranging from 1.8 to 5.6% for other devel-
oped countries (Stout, 1979). It is therefore perhaps unreasonable that
the agricultural industry has to pay world market prices for finite re-
sources such as fossil fuel when it contributes relatively little to their
depletion. The agricultural industry does, however, have obligations,
both social and economic, to examine its current usage of energy and
by processes of systematic analysis to increase its energy efficiency
where possible.
1.2 Systems 5

1.2 SYSTEMS

Any system can be defined as a set of interrelated components ar-


ranged so as to accomplish a desired objective. The system under con-
sideration must have boundaries defining its limits. These boundaries
might be structural in nature, such as a tractor or an enclosed farm.

Energy

~---Energy

System boundary
Fixed mass

Closed system

Mass

Two-flow open system

One-flow open system

Fig. 1.2. Open and closed systems.


6 I. Fundamental Considerations

They might also be apparent rather than structural. The scope of the
problem and the method of analysis will impose boundaries that are
dictated by limitations of analysis time and perhaps by overall objec-
tives. The tractor engine could be viewed as a complete system. In
order to cause the engine to function, energy in the form of fuel must
cross the system boundary, and in order to make use of the available
engine power, the work output must also be transmitted across the
boundary. If we increased the boundary limits, we could then encom-
pass the entire vehicle as a system for analysis. The positioning of the
boundary depends on the objectives of the analysis, and also on the
analytical tools available and our ability to analyze the sets of inter-
acting system components.
Figure 1.2 illustrates what are envisaged as open and closed systems.
The open systems have flows, which can either enter or leave the
system by traveling across the system boundary. Figure 1.3 is a sche-
matic of a much larger system: that of mass, energy, and information
transfers within our world. Two key interacting components are shown,
one representing the natural systems of the environment, and the other
the imposed component due to human establishment of technological
and nontechnological groups in society. Thus the scope of a system is
dictated only by the boundaries of the system and by the goals of the
analysis or investigation.

1.3 ENERGY AND PRINCIPLES OF DYNAMICS

Energy is by definition the capacity of a body to do work. Work is


said to occur when a force moves through a distance. When a force
acts, it produces an effect that, if work is done, results in some rate
of change of momentum. Thus force is equal to mass times acceleration,
or
F = mf
where F is an applied force, m a given mass, and f the acceleration
of the mass. If we consider x and y coordinates, with the mass m at
time t, then with respect to the proposed coordinates

where the dots denote differentiation and


1.3 Energy and Principles of Dynamics 7

------------,

"\
,\
/
/

System boundary
L_

Fig. 1.3. Mass, energy, and information transfers. Solid lines indicate mass/energy
transfers; dashed lines indicate information transfers.

Fy = my
with respect to time. If the point of application of the force moves from
x and y to (x + dx, y + dy), then the work done by the force is
FAx + dx) and
If the point of application moves along a curve between points a,b and
c,d, then the total work done is integrated along the curve as

iX==ac Fx dx + J:==bdFy dy
When a change in state from position 1 to position 2 occurs because
of the application of a force, then the energy of the particle is the work
done by an equal and opposite force to that producing the original
8 I. Fundamental Considerations

change. Forces producing energy must be such that the network is zero
when the particle is brought back to its original state. Some forces
(friction, for example) do no mechanical work on the particle and there-
fore do not produce mechanical energy.

1.4 THE CONSERVATION OF ENERGY

In any given system, the sum of the energy inputs is equal to the
sum of the energy outputs. Not all the energy inputs can go to producing
useful work in the form of a desired output. Losses will occur that are
in reality the transformation of energy into a nonuseful form. The fuel
energy input to a tractor is partly converted to mechanical energy and
also partly to other, nonuseful energy forms such as noise, heat, and
vibration. However, the total energy of the system remains constant.
Energy cannot be created nor can it be destroyed, it can only be
changed in form.
The potential energy of a particle is the energy due to its elevation
above some reference level. Thus for a mass m, the force acting down-
ward is mg, where g is the acceleration of gravity. For an elevation of
distance h, the potential energy (PE) is given by
PE = mgh
The kinetic energy is the work done by a force in overcoming the
inertia of a particle when its speed is changed from Vi to V.
The force is given by F = mx, and the element of work is F?'>x. Thus,
the total kinetic energy KE is

KE = f mx dx f: mx dx
=

For a reference state, Vi can be assumed equal to zero, thus


KE = ! mVZ
The application of the concept of conservation of energy can be dem-
onstrated mathematically by considering a particle with a force acting
such that the particle has potential energy. If the force acting is des-
ignated as (Fx, Fy), the equations of motion are
References 9

or
. di . dy
mx- my-
dx dy
Integrating and adding,

!2 m(i2 + j2) J (Fx dx + Fy dy) + C


Rearranging,

~ mV2 - J (Fx dx + Fy dy) = C

The potential energy of the particle is given by

-J (Fx dx + Fy dy)

The minus sign indicates that, in accordance with the definition of


energy, the force is acting to oppose that acting on the particle. Thus
the potential energy is converted into kinetic energy, and the sum of
kinetic and potential energies at any time is a constant:
PE + KE = const

REFERENCES
Buchele, W. F. (1959). Design and operation of the M.S. U. tandem tractor. Trans. ASAE
2(2), II, 12, 15.
Chamen, W. T. C., Collins, T. S., Hoxey, R. P., and Knight, A. C. (1980). Mechanization
opportunities likely to be provided by engineering in the 21st century. Agric. Engr.
35(3),63-70.
Matthews, J. (1982). The mechanical farm of 2030. Agric. Engr. 37(1), 30-32.
Stout, B. A. (1979). "Energy for World Agriculture." FAO, Rome.
Yahya, R. K., and Goering, C. E. (1977). Some trends in fifty-five years of Nebraska
test data. ASAE paper MC-77-503.
2
Thermodynamics

2.1 THE STEADY-FLOW ENERGY EQUATION

This mathematical statement is an application of the law of conser-


vation of energy to fluid flow in thermodynamic systems.
In addition to potential and kinetic energies, a fluid will possess
internal energy and flow or displacement energy.
The internal energy of a fluid refers to energy it stores as a result of
various processes carried out on the fluid. This energy results from the
internal motion of its atoms and molecules, and their resultant kinetic
effects. The motion activity of these atoms and molecules is reduced
as the temperature of the fluid is decreased, until the internal energy
becomes zero at absolute zero temperature. The symbol for internal
energy is U, with u being used to denote specific internal energy, or
the internal energy per unit mass. Throughout this text the convention
of using an uppercase symbol to represent the thermodynamic property
of any quantity of material and a lowercase symbol to represent a
specific value will be followed.
Flow or displacement energy relates to the energy necessary to cause
a fluid to enter or to leave a system. A given volume of fluid entering
or leaving a system must displace an equal volume immediately in front
in order to permit this entry or exit to take place. If a fluid has a specific
volume VI then in order to enter a system it must displace the same
10
2.1 The Steady-Flow Energy Equation 11

w System

Entry
conditions
P,v,u,C,

Q
x,
Exit
conditions

Fig. 2.1. Hypothetical thermodynamic system.

specific volume VI within the system. Furthermore, if the pressure of


the fluid is PI. then the work done by the fluid on the system is PIVI.
This is the flow or displacement energy received by the system. At the
exit, the same reasoning can be applied to determine the flow or dis-
placement energy lost from the system.
In addition to those energies due to the characteristics of the system
and its working fluid, the system can receive or reject heat and can
perform external work, or work can be performed on the system. The
symbol for energy is Q, and the symbol for work is W.
It is customary to adopt the following algebraic convention to in-
dicate if heat and work are being added to or taken from the system.
Q > 0: heat received by the system
Q < 0: heat rejected from the system
Q = 0: no heat loss or gain
w>o: work done by the system
w<o: work done on the system
W = 0: no external work done on or by the system
Figure 2.1 shows a hypothetical thermodynamic system with inlet
conditions PI, VI, UI, CJ, and XI, and outlet conditions P 2 , V2, U2, C2 ,
12 2. Thermodynamics

and X 2 • Here, P refers to the fluid pressure, u refers to the specific


internal energy of the fluid, v refers to the specific volume of the fluid,
C refers to the fluid velocity, and X is the height of the fluid above
some reference level. The subscripts 1 and 2 refer to the entrance and
exit conditions, respectively.
The law of conservation of energy states: The total energy entering
the system is equal to the total energy leaving the system.
If we consider unit mass, then this relationship can be expressed
mathematically to give the following steady-flow energy equation:
gX I + UI + PIVI + !Ci + Q = gX2 + U2 + P2V2 + !C~ + W
The enthalpy of a fluid is a property that results from considering a
combination of the internal energy and flow energy. Thus specific en-
thalpy is denoted h = U + Pv, where u and v are specific properties.
Total enthalpy is denoted H = U + PV.
By introducing the enthalpy property, the steady-flow energy equa-
tion is modified as follows:
gX I + hI + !Ci + Q = gX2 + h2 + K~ + W (2-1)
Example
A steam turbine has steam entering with the following properties:
velocity = 20 m sec - I

specific enthalpy = 3100 kJ kg - I

The steam leaves the turbine with the following properties:


velocity = 40 m sec - I

specific enthalpy = 2800 kJ kg - I

The steam-flow rate is 100 kg sec - I, and the heat lost from the system
is 30 kJ kg- I. Determine the power output.
Solution. Changes in potential energy are likely to be negligible and
therefore will not be considered. Using Eq. (2-1),
gX I + hI + !Ci + Q = gX2 + h2 + !C~ + W
and neglecting potential energy change,
hI + !Ci + Q = h2 + K~ + W
Heat is lost from the system, and therefore Q is negative.
2.2 The Nonflow Energy Equation 13

The specific work is given by


W = (hI - h 2 ) + ICT - C1) - Q
= 3100 - 2800 20 2 - 40 2 - 30
+ 2 X 103
= 300 - 0.6 - 30
= 269.4 kJ kg-I
The steam flow rate is 100 kg sec-I, and therefore the output power
IS

W = 269.4 x 100 kJ sec-I = 26940 kW


or
W = 26.94 MW

2.2 THE NONFLOW ENERGY EQUATION

When dealing with the analysis of a closed system, some simplifi-


cations can be made to the steady-flow energy equation. For these
cases, the fluid mass in the system remains constant. Thus the flow
terms can be dropped. The nonflow energy equation is
UI + Q = U2 + W (2-2)
2
Note that the flow terms PV and IC2 have been omitted, and
Q = (U2 - ud + W or Q = flu + W (2-3)
For any mass,
Q = flU + W (2-4)
Thus for a nonflow or closed system, the heat transferred through the
system boundary is equal to the change in internal energy of the work-
ing fluid, plus the work added to the system.

Example
A piston in a cylinder compresses a working fluid and in doing so
consumes 80 kJ kg - I of energy. The heat lost to the surroundings is
40 kJ kg-I. Find the change in internal energy.
Solution. This is a closed system since the working fluid does not
14 2. Thermodynamics

cross the system boundary. The nonflow energy equation (2-3) is ap-
plicable:
Q = flu + W
and so
flu Q- W
-40 - (-80) = 40 kJ kg-I
The process has produced a gain in the internal energy of the working
fluid, because of the work done on the system and despite the heat
loss from the system.

2.3 THE FIRST LAW OF THERMODYNAMICS

The law of conservation of energy states that all forms of energy are
mutually convertible. The nonflow energy equation defines this law,
that is, Q = fl U + W. The equation shows that heat and work are
related. However, not all heat energy will be converted into work. As
Eq. (2-4) shows, some energy will be attributable to a change in internal
energy.

2.4 THE SECOND LAW OF THERMODYNAMICS

This law states that heat will not transfer up a temperature gradient
of its own accord. In order for heat transfer from a low-temperature
source to a higher temperature sink to occur, energy must be added
to facilitate the transfer. However, heat will naturally flow from a high-
temperature source to a lower temperature sink. A practical interpre-
tation of this law is that when the temperature ofthe heat energy supply
to a system becomes the same as the system temperature, no heat
exchange can take place. Thus, when a body reaches the temperature
of the coldest surrounding objects, no further unaided extraction of
heat is possible. Engines used for power generation require a source
of fuel that will, when ignited, raise the temperature of the working
substances in the engine above those of the surroundings. The heat
energy generated can then be converted into work.
Whereas the first law of thermodynamics deals with the accounting
of energy by making it equivalent in its various forms, the second law
2.5 The Thermodynamic Engine 15

of thermodynamics recognizes that all forms of energy are not equiv-


alent in terms of their ability to do work. An ideal thermodynamic cycle
is one that is reversible. This is a theoretical concept that cannot be
achieved in practice. However, for the case of the ideal (reversible)
cycle, the available system energy would be conserved. Reversibility
cannot be achieved in practice since it would require an infinite amount
of time to complete the cycle to allow the external temperatures to
vary exactly with the working substance temperatures. Thus for a prac-
tical or real process, the available energy will decrease. This loss in
system available energy represents a reduction in the state of order of
the energy source. This level of order reduction is measured by a ther-
modynamic property called entropy. Thus, for a real process, there is
a gain in system entropy, whereas for the ideal, reversible process,
there is no change in system entropy. The level of entropy varies in-
versely with the temperature ofthe energy form. Gravity, for example,
has zero entropy since there is no associated temperature. At the other
extreme, cosmic microwave radiation is regarded as the ultimate heat
sink since there is no known way to which this energy could be further
degraded. Cosmic microwave radiation has an entropy value per unit
energy of roughly lit relative to unity for sunlight (Dyson, 1971).
Change in entropy was defined in 1864 by Clausius (Tribus and
McIrvine, 1971) as
dS = dQrevlT
where dS represents the change in entropy, Qrev the reversible heat
exchange, and T the absolute temperature at which the heat exchange
occurs.

2.5 THE THERMODYNAMIC ENGINE

Figure 2.2 shows the work and heat energy transfers relative to some
hypothetical thermodynamic engine. The heat energy supplied from
the source (Q) will be converted into work (W) and to other, nonuseful
energy forms, which are represented by (Q - W). The fraction of heat
energy supplied that is converted into useful work is called the thermal
efficiency and is given the symbol 11TH:
11TH = WIQ
16 2. Thermodynamics

Source

1----- W

(Q-W)

Sink

Fig. 2.2. Thermodynamic engine.

This relationship can be expanded, by considering that


W = Q - (Q - W)
i.e., work done = heat received - heat rejected, to show that
T)TH = 1 - (Q - W)/ Q

2.6 GAS LAWS

The work done by a heat engine results from the changes in volume
ofthe working gas contained in the system. If the mass and temperature
of the gas remain constant, then the volume will vary inversely with
pressure. This relationship was first established by the English scientist
Robert Boyle, and is referred to as Boyle's law. Thus,
P1V 1 = P 2 V 2 = PnVn = C
where C is a constant. A plot of P against l!V will be a straight line,
passing through the origin and having slope C.
If a fixed mass of gas has constant pressure, then the relationship
between volume and temperature is such that the volume will vary in
proportion to the absolute temperature of the gas. This is called
Charles' law, after the French scientist Jacques Charles. Thus

where C is a constant.
A plot of volume V against absolute temperature T will give a straight
line of slope C with an intercept at the origin.
2.6 Gas Laws 17

If a gas whose original state is PI, VI, TI is passed to a state P z , V z ,


Tz , then Boyle's and Charles' laws can be combined to give
or PVIT = C
where C is a constant.
If specific volume (i.e. 1 kg) of gas is considered, then the constant
is given the symbol R, where R is the characteristic gas constant. For
any quantity of gas (m kg),
PV = mRT
This is known as the characteristic equation of a perfect gas.
A universal gas constant can be found by considering the application
of Avogadro's law, which states: Equal volumes of different gases at
the same temperature and pressure contain the same number of mol-
ecules.
It follows from this law that, for the same temperature and pressure,
the ratio of molecules contained is equal to the ratio of volumes. Thus
the densities of gases at the same temperature and pressure are pro-
portional to their molecular weight. The mass of substance equal to its
molecular weight is called a mole, symbol mol.
Oxygen has a molecular weight of 32 and, for mass in kilograms,
1 kg mol = 32 kg, and 1 g mol = 32 g
etc. The relative molecular mass of gas
mass of gas M
M = ------=-----
number of moles of gas n

The average relative molecular mass of a gas mixture is given by


M AV = LnMl'Ln
For air, the principal constituents are given in Table 2.1, and
M AV LnMILn
(32 x 21) + (28 x 78.07) + (44 x 0.03) + (40 x 0.9)
100
28.95 kg
(Note, the number of moles = percentage age by volume.)
The gas constant R will differ for each gas if, in the equation PV =
18 2. Thermodynamics

Table 2.1. Constituents of Air

Gas Symbol Mol. wt. Vol. %

Oxygen O2 32 21
Nitrogen N2 28 78.07
Carbon dioxide CO2 44 0.03
Argon A 40 0.9

mRT, V is expressed as volume per unit mass. However, if V is the


volume of 1 mol of gas, then the characteristic gas constant will be the
same for all gases. This constant is called the universal or molar gas
constant. Its value can be calculated for a known gas, say air, as fol-
lows:
MR = RM
where RM is the molar gas constant. For air,
R = 0.287 kJ kg-I K - I and M = 28.95 kg (per mole of air)
Thus
Rrn = 0.287 x 28.95 = 8.313 kJ kg mol- 1 K - 1

The accepted standard value for Rm is 8.3143 kJ kg mol - 1 K - I for


any gas.
The volume of 1 mol of gas can be calculated from
PV = nRMT
For 1 mol, n = 1, and
V = RMTIP
Thus, for any gas at the same temperature and pressure, the volume
can be determined. For example, ifthe pressure is at atmospheric pres-
sure at O°C,
P = 101.33 kN m- 2
T = 273.15 K
V = RMT = 8.3143 x 273.15 22.4 m3 (per 1 kg mol)
P 101.33

2.7 SPECIFIC HEAT

The specific heat of a substance is the amount of heat required to


change a unit mass of substance by 1 degree.
2.8 Gas Processes 19

If C is the specific heat of a substance, then the quantity of heat


transferred Q is
Q = MC(T2 - Td
where T2 - TI is the temperature change of the substance.
Although the value of specific heat of solids and liquids may vary
with temperature, it is usual to use an average value in heat exchange
calculations. The specific heat of a gas, however, will vary depending
on the conditions under which the gas acts.
Two principal specific heat capacities of a gas are the values for
constant pressure and constant volume heat transfer.
The specific heat at constant volume, C v , is a measure of the specific
heat with the volume of the gas held constant. The specific heat at
constant pressure, C p , is a measure of the specific heat with the pres-
sure of the gas held constant.
It can be shown that
Cp - Cv = R
and the ratio of CplC v = -y, the adiabatic index.
For air, average values for specific heat are (Joel 1966)
Cp = 1.005 kJ kg- I and C v = 0.718 kJ kg- I

2.8 GAS PROCESSES

For a closed system, the nonflow energy equation applies. Thus for
the expansion or compression of a gas in a cylinder
Q=tJ.V+W
where Q is the heat added or rejected, tJ. V the change in internal energy,
and W the work done by or added to the system.

2.8.1 The Constant-Volume (Isochoric) Process


Since there is no change in volume, the piston does not change po-
sition and no external work can be done. Thus
W =0 and Q=tJ.V

2.8.2 The Constant-Pressure (Isobaric) Process


The work done by the gas at constant pressure is W = P( V 2 - V I).
Thus
Q = (V2 - VI) + P(V2 - Vd = (V2 + PV2) - (VI + PV I )
= H2 - HI
20 2. Thermodynamics

where H2 - HI is the change in enthalpy of the gas. The heat added


or rejected in a constant pressure process is equal to the change in
enthalpy.

2.8.3 The Constant-Temperature (Isothermal) Process


The changes that result from the isothermal expansion or compres-
sion of a gas are defined by Boyle's law as
PV = C
where C is a constant. The work done is given by the area under the
pressure-volume diagram. Thus
V2

W = ~ PdV
VI

Since for the isothermal process PV = C, then by rearrangement P =


CV- I; substituting gives
V2 dV
W= C ~ -
VI V
or
V2
W = C[ln V]t:~ or W = PV In -
VI
For an isothermal process, Ll U = 0 and from Q = Ll U + W,
Q=W

2.8.4 The Adiabatic Process


A process that is carried out without heat loss or gain is said to be
adiabatic. For the adiabatic process, Q = o.
A process that is carried out in a very small relative time frame can
be considered adiabatic. For example, the bursting of a balloon or a
tire represents a process of rapid expansion of a gas. Because this
occurs in a small period of time, it may be assumed that insufficient
time is available for heat transfer-thus the process is adiabatic. The
adiabatic process is a particular form of the general expression relating
pressure P and volume V for the work done on or by a gas:
pyn = C
where n is the index of expansion or compression and C is a constant.
2.8 Gas Processes 21

The work done is given by the area under the pressure-volume dia-
gram:

For the polytropic process,


or P = cv- n
and so

W = C I
VI
V2
V-ndV

c
- - - v-n+' _ Vl"n+'
-n + 1
Substituting C = Pvn,

W = PIV, - P 2 V2
n - 1

The adiabatic process will obey the polytropic law; however, the index
will have a value that will satisfy the adiabatic requirement of no heat
loss or gain. The adiabatic index 'Y will thus apply:
PV-Y = C or P = CV--Y
The work done is given by
W = P, V, - P 2 V 2
'Y - 1

Also, since Q = d U + W, and for an adiabatic process Q = 0, then


W = -dU
This shows that work and internal energy are related for the adiabatic
process. Internal energy will increase during adiabatic compression and
decrease during adiabatic expansion. The internal energy change d U
is given by

where m is the mass of gas, C v the specific heat at constant volume,


and T2 - T, the temperature change.
22 2. Thermodynamics

p Isothermal PV = C

Adiabatic
pvy=C

v
Fig. 2.3. Camot cycle, pressure-volume diagram.

Since W = - b. U,
mR(T, - T 2 )
- [mC V (T2 - T.)]
'Y - 1

and thus
R
Cv =
'Y - 1

Substituting R Cp - C v gives
Cp - Cv
Cv
'Y - 1

Solving for 'Y yields

'Y =
Cv Cv
Thus the adiabatic index 'Y is given by the ratio of specific heat ca-
pacities at constant pressure (C p ) and constant volume (Cv ). For air,
'Y = 1.4.
References 23

T
Isothermal

Adiabatic
Adiabatic

Isothermal

s
Fig. 2.4. Camot cycle, temperature-entropy diagram.

2.9 THE CARNOT CYCLE

One example of the constant-temperature process is the Carnot


cycle, where heat exchanges occur as isothermal processes. This cycle
was envisaged by the French physicist Sadi Carnot. The pressure-
volume diagram for the complete cycle is shown in Fig. 2.3, with the
appropriate temperature-entropy diagram shown in Fig. 2.4. The Car-
not cycle is a reversible process. It was shown by Carnot, in a paper
published in 1824, that no engine can be more efficient than one that
is thermodynamically reversible. Thermodynamic reversibility means
that the cycle can represent an engine or a heat pump.

REFERENCES

Dyson, F. J. (1971). Energy in the universe. In "Energy and Power" (Scientific American
Editors, editors). W. H. Freeman, San Francisco, CA.
Joel, R. (1966). "Basic Engineering Thermodynamics." Longman, London.
Tribus, M., and Mcirvine, E. C. (1971). Energy and information. In "Energy and Power"
(Scientific American Editors, editors). W. H. Freeman, San Francisco, CA.
3
Power Measurement

3.1 POWER

Power is the rate of doing work. The power output from a tractor
can be measured at several locations. The power developed at the
power takeoff (PTO) shaft is one location that is commonly used in
determining tractor power level. Tests can also be conducted to de-
termine the power output at the drawbar and at the flywheel of the
tractor engine. For a given tractor, output power at the PTO will be
less than that at the engine flywheel because of power losses in driving
the transmission components. Power is also lost at the tire-ground
interface, in driving the engine components, as heat to oil and other
coolants, and to the exhaust. The maximum measurable power ob-
tainable from an engine is that developed inside the cylinders. This
power level is referred to as indicated power (IP).
The power output at the flywheel is referred to as brake power (BP),
and
mechanical efficiency = BP/IP
A measure of transmission efficiency is obtained by comparing PTO
power (PTOP) with BP:
transmission efficiency = PTOP/BP
24
3.2 Torque 25

The power developed at the axles (axle power, AP), can be compared
with the drawbar power (DBP) to give the efficiency of traction, or
tractive efficiency. Thus
tractive efficiency = DBP/AP

3.2 TORQUE

For a rotating shaft, the power P (watts) developed is dependent on


the angular force or torque (T) producing the rotation, and the speed
of rotation:
P = 2'ITNT W
where N is the shaft speed (revolution per second, rps) and T the torque
(Nm).
The torque output required to maintain a given power level is given
by
T = P/2'ITN

Speed

Fig. 3.1. Ideal performance characteristics.


26 3. Power Measurement

Ungoverned range
Power

Engine speed

Rated engine
speed

Fig. 3.2. Typical diesel power output characteristic.

The above equation shows that in order to maintain a constant power


output, the torque output must increase as the shaft speed N is reduced.
Figure 3.1 shows the relationship between torque and shaft speed that
would be required to maintain constant power ouput. An internal-com-
bus ion engine is unable to provide a torque-speed relationship that
enables a constant power output to be maintained. However, the diesel
engines used in tractors have torque-speed characteristics that provide

Torque

----------1-
Torque back-up
__ L_

No-load
speed

Rated
speed

Fig. 3.3. Typical diesel torque output characteristic.


3.3 Indicated Power (IP) 27

a reasonably constant power output over a substantial portion of the


working engine speed range. Typically, the maximum range is from
1300 to 2100 revolutions per minute (rpm) and some tractors are ad-
vertised as having "constant-power" engines, which in reality means
that their torque characteristics provide for a constant power output
over the majority of this range of engine speeds. Figure 3.2 is a typical
plot of power aginst rotational speed (rpm) of the power output shaft.
The speed at which maximum power is obtained is called the rated
engine speed. Above this speed, the engine is operating under governor
control (the governed range) and the power falls as the maximum ro-
tational speed is approached. The torque curve in Fig. 3.3 shows that
torque output rises as the speed falls, until the maximum torque is
reached at typical tractor engine speeds of 1300-1600 rpm. The dif-
ference between the torque output at the rated speed and the maximum
torque is referred to as the torque back-up. This is an indicator of the
"lugging" ability of the tractor, and as a general rule should be at least
10% of the torque at rated speed.

3.3 INDICATED POWER (IP)

The indicated power of an engine is the total power developed in the


cylinders of the engine. The term "indicated" is used because the
power is calculated from the cylinder mean effective pressure, which
was traditionally determined by using a pressure-indicating device that
recorded the pressure changes in the cylinder as the piston moved
through one stroke. The mean effective pressure is an assumed con-
stant-pressure value, which would produce the same power output as
is actually obtained. Figure 3.4 is a typical plot of pressure against
volume from an indicator diagram. The mean pressure of the shaded
area is the mean effective pressure (MEP).
The indicated power (IP) can be calculated by considering the force
exerted on the piston in unit time. Let
piston area (m 2 )
A
Pm MEP (N m- 2 )
N engine speed (rps)
L length of stroke (m)
f factor for the number of power strokes per revolution
of the engine crankshaft
28 3. Power Measurement

Pressure

Swept I
volume ----,
TDC BDC
Volume

Fig. 3.4. Indicator diagram. [TDC and BDC are, respectively, top and bottom dead
center, the upper- and lowermost positions reached by the piston.]

Then for a four-stroke engine, there is one power stroke every other
revolution of the crankshaft, i.e.,
f = 112
For a two-stroke engine, there is one power stroke every revolution,
and so
f = 1
The mean force F m acting on the piston is
Fm = PmA
The work done is
W = PmLA
and considering the number of working strokes, the indicated power
is given by
IP = PmLANf
3.3 Indicated Power (IP) 29

The brake power (BP) is less than the indicated power (lP) because
of losses to friction and to driving auxiliaries such as the alternator and
oil pump. The ratio between BP and IP was given in Section 3.1 as the
mechanical efficiency. The difference between BP and IP is often
termed the friction power FP:
IP - BP = FP
An alternative method of determining IP is to use the Morse test. This
approach allows the IP to be determined by using the above relationship
between IP, BP, and FP. The brake power (BP) of the engine is de-
termined at a suitable load and operating speed. One cylinder of the
engine is isolated so that it no longer performs any useful function.
This can be accomplished by removing a spark plug lead in the case
of a spark ignition engine or by slackening the appropriate injector fuel
supply line in the case of a compression ignition engine. The load is
then adjusted so as to restore the engine speed to the setting for all
cylinders functioning. The BP output is then recorded. Let BP be the
brake power with all cylinders functioning, ii, i2 , • • • , in the indicated
power of each cylinder, and II, I2, ... , in the friction power of each
cylinder. Then the total IP and BP are given by
n
IP = BP - ~ in (3-1)

n
IP(n-1) = BP(n-l) + ~ in (3-2)
1

Subtracting Eq. (3-2) from Eq. (3-1) gives


BP - BP(n-1) =
where i is the indicated power of the isolated cylinder. This procedure
can be repeated by then isolating other cylinders in turn and adjusting
the load to maintain the desired operating speed. Values for IP for each
cylinder over the working engine speed range can be determined by
repeating the above steps at chosen levels of engine speed.
An assumption of the Morse method is that the friction power (FP)
remains the same for the working state and the isolated or "free-wheel"
condition. This will lead to some inaccuracies, but it is generally agreed
that it gives a reasonable approximation.
30 3. Power Measurement

Example
Let
Pm = 620 k Nm- 2 , N = 30 sec-I, A = 0.12 m2 , L = 0.15 m
Determine the IP of this four-stroke, four-cylinder engine.
Solution. For n the number of cylinders
IP = PmLANfn = 620 x 0.15 x 0.12 x 30 x 0.5 x 4
= 67.0 kW
Under test, the engine develops a brake power output BP of 54 kW
at a speed of 35 rps (revolutions per second).
Thus the mechanical efficiency is

BP = 54 = 081 or 81%
IP 67 .
Each cylinder was isolated in turn and the load adjusted so as to main-
tain the set engine speed of 30 rps (1800 rpm).
The following results (kW) were obtained:

Cylinder isolated BP(n-1) i (BP - BP(n-l))

I 39 54 - 39 = 15
2 37 54 - 37 = 17
3 36 54 - 36 = 18
4 37 54 - 37 = 17

and IP = 67 kW. The Morse test method thus gives a value for IP that
is in agreement with that obtained by considering the MEP and the
engine dimensions.

3.4 ENERGY BALANCE

In the previous sections of this chapter we defined the terms indi-


cated power and brake power, and showed that their ratio expresses
a value for the mechanical efficiency of the engine. The difference
between indicated power and brake power represents a power loss,
which was defined as friction power. It is interesting to compare the
proportion of input energy that goes to performing useful work with
the proportion of the input energy that goes to nonuseful sinks.
3.4 Energy Balance 31

Table 3.1. Calorific Values at ISS C (kJ kg-l)a

Fuel c:; c:
Gasoline 43,800 46,800
Diesel 42,700 45,400
Kerosene 43,300 46,200

a From Bacon (1980).

The energy input results from the rate of fuel supplied. This energy
input is given by

where in is the mass flow rate (kg sec - I) and C v the calorific value (J
kg-I). Two values for the calorific (or heating) value of a fuel con-
taining hydrogen can be obtained. The higher (or gross) calorific value
C:: is obtained when the water formed by combustion is condensed.
The lower (or net) calorific value C:; is obtained when the water formed
by combustion is in the vapor state. The difference between the
C:: and the C:; is the latent heat of vaporization of water at the given
temperature. In practice it is likely that the water vapor is not con-
densed as a result of fuel combustion in an internal combustion engine,
so that it may be considered as leaving the system along with the ex-
haust gases. Thus C:; is more appropriate for determining the fuel
energy input to an engine. Calorific values are given in Table 3.1.
The BP can be expressed in terms of energy units as kJ sec - I. The
energy that goes to the coolant water can be determined, if the mass
flow rate and temperature rise of the coolant are known, by
in w C(t2 - tl) J sec-I
where inw is the mass flow rate of coolant water (kg sec - I), C the
specific heat capacity of water (=4.18 kJ kg-I K- 1 ), and t2 - tl the
temperature rise CC) of the coolant water.
Energy to the exhaust is determined in a similar manner if the mass
flow rate of exhaust gas in known in addition to the specific heat ca-
pacity of the exhaust gas and the difference between the exhaust gas
and ambient temperatures (tE - tA):
inECE(tE - tA) J sec-I
32 3. Power Measurement

Radiation and other losses

Exhaust heat
80

-+-- Cooling water heat

Friction losses

Net effective work (BP)

o L -_ _ _- - '_ _ _ _..J
o 50 100
Percentage of load

Fig. 3.5. Heat balance of a typical diesel engine. [Reproduced from Liljedahl et al.
(1979) with the kind permission of John Wiley & Sons, copyright 1979.]

It is difficult, however, to get a reliable estimate of CEo Thus, a more


useful approach is to use an exhaust gas heat exchanger (exhaust gas
calorimeter) .
The calorimeter usually consists of a counterflow heat exchanger,
with water as the heat exchange medium. Thus, if in is the mass flow
rate of water (kg sec-I), tl and t2 the inlet and outlet temperature of
the water eC), respectively, C the specific heat capacity of water ( =
4.18 kJ kg-I K -I), and the energy gained by the water equals the
energy lost by the exhaust, then the energy in the exhaust is
inC(t2 - td(tE - tA)I(tA - t3) J sec-I
wheretE is the exhaust gas temperature entering the calorimeter eC),
tA the ambient temperature eC), and t3 the exhaust gas temperature
leaving the calorimeter eC).
The calorimeter method can also be used to measure the energy given
up to the lubricating oil. The heat lost to the surrounding air is difficult
to measure, but it can be determined by considering that all the heat
that is unaccounted for is dissipated in this manner. Thus, the heat loss
to the surrounding air is given by
fuel energy - (BP energy + coolant energy + exhaust energy)
3.5 Measurement of Fuel Flow 33

Figure 3.5 shows how the fuel energy is dissipated to the principal
energy sinks that have been considered in the above analysis. This
energy balance shows that, at full load, approximately 28% of the fuel
energy goes to the cooling water, about 26% is lost to the exhaust,
about 35% is converted into useful work, and the remainder, approx-
imately 11%, is lost to friction and radiation (Liljedahl et al., 1979).

3.5 MEASUREMENT OF FUEL FLOW

Fuel flow measurements are based on either mass or volumetric


changes. The advantage of the mass measuring system is that it is
temperature independent. For this reason it provides a more accurate
method of measuring fuel consumption and is usually employed when
engines are tested on stationary dynamometers. Volumetric measure-
ment is somewhat simpler to arrange, but the viscosity of diesel fuel
varies considerably with temperature. To obtain an accurate measure-
ment of fuel consumption by volumetric methods it is necessary to
apply fuel temperature corrections for the effects of viscosity and den-
sity of the fuel. The fuel consumption can be compared with the power
output at different loading conditions. This provides a measure of the
efficiency with which the fuel is being converted into useful work, and
is termed the specific fuel consumption:
SFC = fuel consumption/power out
The units of SFC are kg kW- I hr- I for mass flow measurement of
fuel and liters kW- 1 hr- I for volumetric measurement of fuel flow.
The accurate measurement of fuel flow on a tractor that is in field
work is somewhat more difficult. Mass-based systems are not suited
to a mobile vehicle, and thus volumetric-based systems are commonly
used. A number of volumetric systems for in-field measurement of
tractor fuel use have been proposed. Reid (1979) used two plexiglass
graduated fuel cylinders, which could be used individually or together
depending on fuel demand. Tomkins and Carpenter (1980) used a sim-
ilar system, as did Beppler and Shaw (1980). Others, such as Frisby
and Summers (1970), Lin et al. (1980), and Bedri et al. (1981), used
turbine flow meters inserted in the fuel line. Bedri et al. used heat
exchangers to cool the fuel recirculated from the injectors and fuel
pump so as to reduce temperature effects. Positive displacement flow
meters have been used by Grevis-James et al. (1981), Smith et al.
34 3. Power Measurement

(1981), and Clark and Gillespie (1979). An advantage of this approach


is that the output from the positive-displacement flow meter is unlikely
to be affected by viscosity changes.
Wilkes (1982, personal communication, London Univ., Kent) estab-
lished a relationship between the metering pin position of the rotary
injection pump and fuel consumption. This approach does not require
any modification to the existing fuel lines, but it does require monitoring
of fuel temperature. Some error will be present because of leak-off
from the injectors.
An approach that has been adopted by a number of investigators is
simply to refill the fuel tank after running the tractor in the field over
a prescribed distance. Measuring the volume of fuel required to refill
the tank provides a realistic estimate of the average fuel use over that
distance. This method is simple and fairly reliable and does not interfere
with the existing fuel lines, which is an important consideration if
farmer-owned tractors are being field tested. A simple unit capable of
measuring and recording tractor fuel consumption in the field without
interfering with existing fuel lines is required. A number of options
exist, such as level indicators for the fuel tank using ultrasonic trans-
ducers to monitor fuel height, or secondary tanks arranged so as con-
tinually to maintain a preset level in the main fuel tank.

3.6 MEASUREMENT OF AIR FLOW

The measurement of air flow is important in engine testing since it


enables the volumetric efficiency to be determined. The volumetric
efficiency of a normally aspirated reciprocating piston-type engine is
the ratio of the volume caused to flow into the cylinder to the swept
cylinder volume.
The usual method of measuring air flow to an engine is to draw the
intake air from a surge tank and to measure the resultant flow of air
into the surge tank by a calibrated orifice.
The flow through an orifice is given by the expression
actual discharge = Cd x theoretical discharge
where Cd is the coefficient of discharge for the orifice. Thus
Q = [Cd AdV(1 - m 2 )]Y2(Pl - P 2 )1-y
where Q is the actual discharge, m = A dA2' A 1 the area of the orifice,
3.7 Tractor Performance 35

PI - P 2 the pressure difference across the orifice, 'Y the specific mass
of air, and g the acceleration due to gravity.
For a sharp-edged orifice, and for dID ratios in the range 0.2-1.0,
Cd = 0.601 (Spink, 1967). For air, 'Y = 1.131 kg m- 3 . Thus
Q = [0.799AdVO - m 2 )]V(P I - P2)
A fuller account of fluid-metering principles can be obtained by con-
sulting the many texts dealing with fluid mechanics, such as Spink
(967) and Barna (971).

3.7 TRACTOR PERFORMANCE

The output power of a tractor is usually determined at the PTO shaft


and the drawbar. The PTO shaft may be driven at either 540 or 1000
rpm. The 1000-rpm drive is available on the larger models of tractor.
Since the power output P from a rotating shaft is given by
P = 2'ITNT
where P is the power (W), N the shaft speed (rps), and T the shaft
torque (Nm), then increasing the PTO speed from 540 to 1000 rpm will
reduce the shaft torque required to give a desired power output.
The PTO power output can be measured on a dynamometer of the
type used to measure BP. Special-purpose PTO power dynamometers
are, however, available and widely used, not only in colleges and uni-
versities, but also in service stations. These dynamometers can be used
to provide values for PTO power at the rated engine speed, and can
also be used to load down the engine over the normal working engine
rpm range. The PTO power and torque curves can thus be determined.
It is also possible to conduct the Morse test to determine the indicated
power of each cylinder.
The drawbar power of a tractor is measured by using a drawbar
dynamometer to measure drawbar pull, and some means of measuring
the tractor speed:
drawbar power = FD V
where FD is the drawbar pull (kN) and Vis speed (m sec-I). The torque
available from the tractor engine must overcome transmission losses
and rolling resistance in providing a drawbar pull.
36 3. Power Measurement

I Maximumsustained
I drawbar pull

Drawbarpull

Fig. 3.6. Wheelslip vs. drawbar pull.

3.8 DRAWBAR PULL (FD )

The pull developed at the drawbar is dependent on the tractive ef-


fort that can be developed at the tire-ground interface. It can be as-
sumed that the drawbar pull is proportional to the vertical load on the
tractor driving wheels, and that for a constant vertical load, the wheel-
slip will increase exponentially with drawbar pull up to some value at
which the maximum sustained pull is obtained. This effect is illustrated
in Fig. 3.6.
A definition of drawbar pull that is provided by ASAE (S296) is that
it is the force in the direction of travel produced by the vehicle at the
drawbar.
Drawbar pull FD is given by
FD = H - pW

where H is the tractive effort, p the coefficient of rolling resistance,


and W the weight on the tractor driving axle.
The tractive effort H can be obtained from

H = T/r
where T is the driving wheel input torque and r the rolling radius of
the driving wheels.
The coefficient of rolling resistance p is given by
p = R/W
3.8 Drawbar Pull (FD ) 37

I Set speed Engine

1[\
v..... Engine control
system (governor)
~
Speed Torque Speed
change

\I

Transmission / Transmission
I' Gear control
selection
I, 1

Torque Load

\11 Tractive
effort

Tire-ground Speed Tractor


/
..... ~

'I
1[\

Vertical Draft
load

.....
",
Implement Implement control
", system
......
Lift/lower

Fig. 3.7. Dynamic performance of tractor and implement.

where R is the motion or rolling resistance. The weight Won the tractor
driving axle is composed of components due to the tractor weight, the
implement weight, and the vertical soil forces acting on a soil-engaging
implement.
The draft force on an implement can be defined as the horizontal
component of forces resisting the forward motion of an implement. The
difference between the drawbar pull and the draft forces gives the force
that is available to accelerate the tractor and implement combination.
Thus, the forward motion of a tractor and implement can be described
by
~ Table 3.2. Nebraska Tractor Test 1447-Allis Chalmers 8030 Powershift 12-Speed Diesel

POWER TAKE-OFF PERFORMANCE Department of Agricultural Engineering


Power Crank Fuel Consum ption Temperature of iC) Dates of Test: August 3 I-September 13, 1982
Hp shaft Air Air Barometer
(oW) speed gallhr lblhp.hr Hp.hr/gal Cooling wet dry inch Hg
rpm (Uh) (ltg/oW. h) (kW.hll) medium bulb bulb (kPa)
Manufacturer: ALLIS CHALMERS CORPORA-
TION, P.O. Box 512, Milwaukee, Wisconsin
MAXIMUM POWER AND FUEL CONSUMPTION 53201
Rated Engine Speed-Two Hours (PTO Speed-l02l rpm) FUEL, OIL AND TIME: Fuel No. 2 Diesel
133.75 2500 8.882 0.463 15.06 188 66 75 29.030
(99.74) (33.622) (0.282) (2.967) (86.5) (18.8) (239) (98.030)
Cetane No. 46.6 (rating taken from oil company's
inspection data) Specific gravity converted to 60°/
Standard Power Take-off Speed (1000 rpm)-One Hour
134.42 15.25 190 75 29.025
60° (15°/15°) 0.8376 Fuel weight 6.974 Ibs/gal
2252 8.814 0.457 66
(100.24) (33.365) (0.278) (3.004) (87.8) (18.8) (23.9) (98.013) (0.836 kg/l) Oil SAE 15W-40 API service classi-
fication SF-CD To motor 4.000 gal (15.1411)
VARYING POWER AND FUEL CONSUMPTION-Two Hours
Drained from motor 3.731 gal (14.1221) Trans-
119.84 2423 8.337 0.485 14.38 187 65 75 mission and final drive lubricant Allis Chalmers
(89.36) (31.559) (0.295) (2.832) (86.1) (18.3) (23.9)
Power Fluid 821 Total time engine was operated
0.00 2583 3.037 171 66 78 36.0 hours.
(0.00) (11.496) (77.5) (18.9) (253)

61.84 2500 5.657 0.638 10.93 180 65 74 ENGINE: Make Allis Chalmers Diesel Type
(46.11) (21.414) (0.388) (2.153) (82.2) (18.3) (23.3)
six cylinder vertical with turbocharger Serial No.
133.77 2300 8.857 0.462 15.10 188 65 74 70-2290 I Crankshaft lengthwise Rated rpm
(99.75) (33.527) (0.281) (2.975) (86.7) (18.3) (23.6)
2300 Bore and stroke 4.25" x 5.00" (107.9 mm x
31.34 2534 4.315 0.960 7.26 175 66 76 127.0 mm) Compression ratio 15.5 to I Displace-
(23.37) (16.334) (0.584) (1.431) (79.4) (18.9) (24.7)
ment 426 cu in (6981 ml) Starting system 12 volt
91.43 2464 6.995 0.534 13.07 182 65 75 Lubrication pressure Air cleaner two paper ele-
(68.18) (26.479) (0.325) (2.575) (83.6) (18.3) (23.9)
ments Oil filter two full flow cartridges Oil cool·
Av 73.04 2467 6.199 0.592 11.78 181 65 75 29.010 er engine coolant heat exchanger for crankcase
Av (54.47) (23.466) (0.360) (2.321) (82.6) (18.5) (24.1) (97.962)
oil, radiator for hydraulic and transmission oil
Fuel filter two paper cartridges Muffler under-
hood Exhaust vertical Cooling medium temper-
ature control two thermostats.
DRAWBAR PERFORMANCE CHASSIS: Type standard with duals Serial
Power Drawbar Speed Crank- Slip Fuel Consumption Temp. 'F("C) No. S0305 1422 Tread width rear 59" (1499 mm)
Hp pull mph shaft % gal/hr lblhp.hr Hp.hr/gal Cool- Air Air Barom. to 123" (3124 mm) front 62" (1574 mm) to 90" (2286
(kW) lb. (kmlh) speed (Uh) (kg/kW.h) (kW.hll) ing wet dry inch Hg
(W) rpm med bulb bulb (kE'a)
mm) Wheel base 106" (2692 mm) Center of grav-
ity (without operator or ballast, with minimum
Maximum Available Powel'-Two Hours 7th (2F) Gear
115.22 6816 6.34 2299 4.67 8.908 0.539 12.93 189 71 75 29.045 tread, with fuel tank filled and tractor serviced for
(85.92) (30.32) (10.20) (33.721) (0.328) (2548) (86.9) (21.4) (23.6) (98.081) operation) Horizontal distance forward from cen-
75% of Pull at Maximum Powel'-Ten Hours 7th (2F) Gear ter-line of rear wheels 2S.4" (721 mm) Vertical dis-
93.86 5165 6.81 2440 3.52 7.904 0.587 11.87 186 72 75 28.872 talice above roadway 41.2" (1047 mm) Horizontal
(69.99) (22.98) (10.97) (29.921) (0.357) (2.339) (85.3) (21.9) (24.1) (97496) distance from center of rear wheel tread 0" (0 mm)
50% of Pull at Maximum Powel'-Two Hours 7th (2F) Gear to the rightlleft Hydraulic control system direct
64.17 3443 6.99 2470 2.18 6.417 0.697 10.00 185 73 79 29.005 engine drive Transmission selective gear fixed
(47.85) (15.32) (11.25) (24.290) (0.424) (1.970) (85.0) (22.8) (26.1) (97.946) ratio with partial (6) range operator controlled
50% of Pull at Reduced Engine Speed-Two Hours 9th (3F) Gear powershift Advertised speeds mph (kmlh) first
64.16 3443 6.99 1792 2.18 5.431 0.590 11.81 189 76 82 28.980 I.S (2.9) second 2.5 (4.0) third 3.5 (5.6) fourth 4.4
(47.84) , (15.32) (11.25) (20.558) (0.359) (2.327) (87.2) (24.2) (27.5) (97.86J)
(7.1) fifth 4.9 (7.9) sixth 5.4 (8.7) seventh 6.S
MAXIMUM POWER IN SELECTED GEARS (1l.0) eighth 6.S (1l.0) ninth 9.3 (15.0) tenth 1l.S
(19.0) eleventh 14.6 (23.5) twelfth IS.4 (30.0) re-
103.58 13348 2.91 2315 14.93 3rd (35) Gear 185 66 71 28.680
(77.24) (59.37) (4:68) (85.0) (18.9) (21.7) (96.848) verse 2.9 (4.7),7.9 (12.7) Clutch multiple wet disc
112.81 10791 3.92 2301 8.47 4th (45) Gear 187 68 70 29.070
hydraulically power actuated and operated by foot
(U13) (48.00) (6.31) (85.8) (20.0) (21.1) (98.165) pedal Brakes multiple wet disc hydraulically pow-
116.52 9743 4.48 2302 7.03 5th (IF) Gear 187 68 70 29.070
er actuated and operated by two foot pedals which
(86.89) (43.34) (7.22) (85.8) (20.0) (21.1) (98.165) can be locked together Steering hydrostatic
110,36 8273 5.00 2300 5.85 6th (55) Gear 186 68 69 29.070
Turning radius (on concrete surface with brake
(82.29) (36.80) (8.05) (85.6) (20.0) (20.6) (98.165) applied) right 149.7" (3.80 m) left 149.7" (3.80 m)
116.65 6886 6.35 2299 4.63 7th (2F) Gear 187 68 69 29.070
(on concrete surface without brake) right 167.5"
(86.99) (30.63) (10.22) (85.8) (20.0) (20.6) (98.165) (4.25 m) left 16S.4" (4.28 m) Turning space dia-
111.64 6526 6.42 2302 4.32 8th (65) Gear 186 67 68 29.070 meter (on concrete surface with brake applied)
(83.25) (29.03) (10.32) (85.6) (19.4) (20.0) (98.165) right 320.0" (8.13 m) left 320.0" (8.13 m) (on con-
114.69 4837 8.89 2301 3.07 9th (3F) Gear 186 67 68 29.070
crete surface without brake) right 355.5" (9.03 m)
(85.53) (21.51) (14.31) (85.6) (19.4) (20.0) (98.165) left 357.0" (9.07 m) Power take-off 540 rpm at
2230 engine rpm and 1000 rpm at 2252 engine
Ul
rpm.
CD

(continued)
Table 3.2. (continued)

LUGGING ABILITY IN 7th (2F) GEAR REPAIRS and ADJUSTMENTS: No repairs or


Crankshaft Speed rpm 2299 2073 1847 1614 1381 1153 adjustments.
Pull-lbs 6886 7846 8252 8240 8040 7161
(W) (30.63) (34.90) (36.71) (36.65) (35.76) (31.85)
REMARKS: All test results were determined
Increase in Pull % 0 14 20 20 17 4 from observed data obtained in accordance with
Power-Hp 116.65 118.91 110.97 96.83 80.92 60.55 SAE and ASAE test code or official Nebraska test
(kW) (86.99) (88.67) (82.75) (72.21) (60.35) (4516) procedure. For the maximum power tests, the fuel
Speed-Mph 6.35 5.68 5.04 4.41 3.77 3.17 temperature at the injection pump return was
(kmlh) (10.22) (9.15) (8.12) (7.09) (6.07) (510) maintained at 178°F (81.3°C). Seven gears were
Slip % 4.63 5.32 5.62 5.77 5.62 4.86 chosen between 15% slip and 10 mph (16.1 kmlh).
TRACTOR SOUND LEVEL WITH CAB dB(A) We, the undersigned, certify that this is a true
Maximum Available Power-Two Hours 78.0 and correct report of official Tractor Test No.
75% of Pull at Maximum Power-Ten Hours 79.5 1447.
50% of Pull at Maximum Power-Two Hours 79.5 LOUIS I. LEVITICUS
50% of Pull at Reduced Engine Speed-Two Hours 76.0 Engineer-in-Charge
Bystander in 12th (6F) gear 90.0
K. VON BARGEN
TIRES, BALLAST AND WEIGHT With Ballast Without Ballast W. E. SPLINTER
Rear Tires -No., size. ply & psi (JcPa) Inner Two 18.4-38; 8; 14 (95)InnerTwo 18.4-38; 8; 14 (95)
Outer Two 18.4-38; 6; 14 (95)OuterTwo 18.4-38; 6; 14 (95) L. L. BASHFORD
Ballast -Liquid (each) None None Board of Tractor Test Engineers
-Cast Iron (each inner) 910lb (413 kg) None
Front Tires -No .• ,ize, ply & p,i (RJ'a) Two 11.00-16; 8; 40 (275) Two 11.00-16; 8; 40 (275)
Ballast -Liquid (each) None None
-Cast Iron (each) 133 Ib (60 kg) None
Height of Drawhar 20.5 in (520 mm) 20.5 In (520 mm)
Static Weight with Operator-Rear 12820 Ib (5815 kg) 11000 Ib (4990 kg)
Front 4280 Ib (1942 kg) 40151b (1821 kg)
Total moo Ib (7757 kg) 15015 Ib (6811 kg)
3.9 Nebraska Tractor Tests 41

FD - D

where D is the implement draft, Wi the implement weight component,


W t the tractor weight component, V the tractor forward speed, and t
the time.
The dynamic performance of the tractor-implement combination is
affected by the engine torque-speed characteristics, the transmission
efficiency, the efficiency with which tractive effort can be utilized as
drawbar pull through the tire-ground interface, and the characteristics
of the implement depth or draft control system. A block diagram of
these relationships is given in Fig. 3.7 (after Dwyer, 1970).
The required engine torque is given by,

I dN Hr
T= - - +-
g dt T]n

where T is the engine torque, I the polar moment of inertia of the


rotating parts, N the engine rotational speed, T] the transmission effi-
ciency, and n the overall transmission ratio.

3.9 NEBRASKA TRACTOR TESTS

The Nebraska tractor tests have, since 1920, provided a unique basis
for the evaluation of the performance of agricultural tractors. The bill
requiring tractors sold in the state of Nebraska to be tested by a board
ofthree test engineers was enacted on July 15, 1919. This bill had been
introduced by a Nebraska farmer who was concerned about the lack
of knowledge pertaining to tractor performance in those early days of
their development.
The aim of the Nebraska test has been to provide a procedure that
allows comparisons to be made between test results. As a consequence
the drawbar tests are all carried out with the tractor ballasted and
operating on a concrete test track for a specified duration and at a pre-
established percentage of maximum engine power. The draw bar per-
formance tests are in two parts: varying power (at 100, 75, and 50%
of maximum power) and maximum power in selected gears.
42 3. Power Measurement

The lugging ability of a tractor is demonstrated by initially operating


at maximum power and then incrementally increasing the drawbar load
so as to cause a decrease in forward speed.
The PTO performance is determined at maximum power and under
six different power output levels. The maximum PTO output is deter-
mined with the engine operating at rated speed for a duration of 2 hr
and a 1-hr test is used to determine the power output from the PTO
when it is operating at the standard PTO speed.
Sound measurements are taken during each of the varying power
tests. The dB(A) sound level is taken with the pickup microphone lo-
cated next to the right ear of the tractor operator. Bystander sound
readings are taken at a distance of 7.6 m from the line of travel of the
tractor. Table 3.2 is a copy of Nebraska Tractor Test 1447, which is
reproduced with the kind permission of the Engineer-in-Charge (Ne-
braska Tractor Tests), and Allis Chalmers Corporation.

REFERENCES
Bacon, D. H. (1980). Fluid mechanics. In "Mechanical Engineer's Reference Book,"
Section 2 (A. Parrish, ed.). Butterworths, London.
Barna, P. S. (1971). "Fluid Mechanics for Engineers," 3rd Ed. Butterworths, London.
Bedri, A. R., Marley, S. J., and Buchele, W. F. (1981). Tractor performance monitor
based on a single chip microprocessor. ASAE Paper No. 81-1578. Am. Soc. Agric.
Eng., St. Joseph, MI.
Beppler, D. C., and Shaw. M. D. (1980). Tractor instrumentation package for field energy
studies. ASAE Paper No. 80-1522. Am. Soc. Agric. Eng., St. Joseph, MI.
Clark, J. H., and Gillespie, J. R. (1979). Development of a tractor performance meter.
ASAE Paper No. 79-1616. Am. Soc. Agric. Eng., St. Joseph, MI.
Dwyer, M. J. (1970). The dynamic performance of tractor-implement combinations.
Proc. Instr. Mech. Engr. 184, Part 3Q, 68-82.
Frisby, J. C., and Summers, J. D. (1979). Energy-related data for selected implements.
Trans. ASAE 22,1010-1011.
Grevis-James, I. W., DeVoe, D. R., Bloome, P. D., and Batchelder, D. G. (1981). Micro-
computer based data acquisition system for tractors. ASAE Paper No. 81-1578. Am.
Soc. Agric. Eng., St. Joseph, MI.
Liljedahl, J. B., Carleton, W. M., Turnquist, P. K., and Smith, D. W. (1979). "Tractors
and Their Power Units," 3rd Ed. Wiley, New York.
Lin, T. W., Clark, R. L., and Adsit, A. H. (1980). A microprocessor based data acqui-
sition system to measure performance ofa small four-wheel drive tractor. ASAE Paper
No. 80-5525. Am. Soc. Agric. Eng., St. Joseph, MI.
Reid, J. T. (1979). A system for measuring tractor fuel use on small plots. Trans. ASAE
22, 57-59, 62.
Smith, L. A., Barker, G. L., and Colwick, R. F. (1981). Instrumentation used to monitor
References 43

energy requirements for agriculture and field operations. ASAE Paper No. 81-1043.
Am. Soc. Agric. Eng., St. Joseph, MI.
Spink, L. K. (1967). "Principles and Practice of Flow Meter Engineering," 9th Ed.
Foxboro Co., Foxboro, MA.
Tomkins, F. D., and Carpenter, R. G. (1980). Fuel consumption during selected tillage
operations. Tenn. Farm Home Sci. 114, Agricultural Experiment Station, Univ. of
Tennessee.
Wilkes, J. M. (1982). Personal communication. Wye College, London University, Kent,
England.
4
Traction

4.1 TRACTION IN SOILS

Soils, like metals, can behave both elastically and plastically. Elastic
deformation refers to the ability of the deformed material to return to
its original dimensions. Plastic deformation refers to a condition of
permanent deformation. For a soil in the elastic condition, a given
applied force causes a known deformation. On removal of the force,
recovery takes place.
If, however, the force is continually increased, a loading condition
will occur that will cause the soil to deform permanently, i.e., it will
behave plastically. The onset of this plastic condition is generally con-
sidered to be induced by shear failure, i.e., the sliding of one particle
over another. In this case, the ability of a particular soil to support a
given load before a permanent change to the soil structure occurs is
called the shear strength of the soil.

4.2 SHEAR STRENGTH

Granular materials such as soils exhibit cohesive and frictional prop-


erties. Cohesion is the bonding together of soil elements irrespective
of the type of applied load. Pure clay fits this category; dry sand, on
44
4.2 Shear Strength 45

Bulk
shear
strength

Practical Soil moisture


limit content
of soil
moisture
content

Fig. 4.1. Soil bulk shear strength vs. moisture content.

the other hand, exhibits a frictional resistance to shear loads in that


the resistance to shear increases with applied load.
Most soils exhibit a combination of cohesive and frictional proper-
ties. The water content in clay soil has a strong influence on shear
strength: the higher the water content, the lower the shear strength.
An applied load will result in a wedge-shaped cone below the tire,
which will cause deformation and displacement of adjacent soil par-
ticles. If the resultant shear stresses are greater then the soil can sus-
tain, sinkage will occur, increasing the surface area. This will continue
until the soil is able to support the tire and load.
Two forms of shear strength may be considered:
1. Bulk shear strength: the resistance offered to movement by a
relatively large volume of soil aggregates.
2. Clod shear strength: the resistance offered by the individual clod
or aggregate.
The main factors that influence shear strength are
1. moisture content
2. packing density and particle size
3. organic matter content.
Figure 4.1 shows the changes in bulk shear strength with changes in
moisture content. The moisture content is also related to the upper and
lower plastic limits.
46 4. Traction

4.3 PLASTIC LIMIT

As the drying process continues, the plastic state reaches a consis-


tency at which the soil ceases to behave as a plastic and begins to break
apart and crumble.
The increase in shear strength with decreasing moisture content from
the upper to the lower plastic limit is clearly seen. The lower plastic
limit represents the maximum moisture content where a farmer can
break clods during seedbed preparations without causing structural
damage. It is a condition frequently accepted as the upper moisture
limit for working soils in agriculture.
At high moisture contents, clods are very weak, and susceptible to
deformation. The higher the organic matter content, the stronger the
aggregates. Aggregates produced on fine sand and silt soils tend to be
very weak.

4.4 COULOMB AND MICKLETHWAITE EQUATIONS

The failure of an agricultural soil can be described by Coulomb's


equation:
T = C + IT tan <!>
where T is the shear stress of the material, c the cohesive property of
the material, IT the normal stress on the sheared surface, and <!> the
angle of internal shearing resistance of the material.
For saturated clay, the cohesion is independent of the applied normal
load, and so
T = C

Sandy soils, however, have little cohesion but have larger values for
the angle of internal shearing resistance. Thus for a sandy soil,
T = IT tan <!>
An agricultural soil is composed of both sand and clays, and therefore
has properties intermediate between those for sand and clay alone. The
typical agricultural soil has therefore both cohesive and frictional prop-
erties. Figures 4.2 and 4.3 show the shear versus normal stress char-
acteristics for clay and sand soils, and for agricultural soils, respec-
tively.
4.4 Coulomb and Micklethwaite Equations 47

Cohesive soil

u Fig. 4.2. Shear stress or vs. normal


Frictional soil stress (T for cohesive and frictional soils.

Micklethwaite (1944), as cited by Reece (1966), proposed the fol-


lowing modification to the Coulomb equation.
Let A be the contact area; then
TA = cA + uA tan <l>
1fT = Tmax , then TA is the maximum thrust, or Hmax. The normal stress
u = W/A, and so

Hmax = cA + W tan <l>

This expression is usually referred to as the Micklethwaite equation.


The pressure under a rigid wheel on frictionless soils at small sin-
kages was given by Reece (1966) in citing Uffelmann (1961) as
pic = S.7
where p is the pressure. The sinkage z necessary to support the wheel
for radial pressures given by the above equation is
z = W 2 /(S.7c?b 2 d
48 4. Traction

T T=c+atan¢

Cohesion

a
Agricultural soil

Fig. 4.3. Shear stress T VS. normal stress C7 for typical agricultural soil.

where W is the vertical load on the wheel, b the width of the wheel,
and d the wheel diameter. The rolling resistance can be considered as
principally due to the work done in forming the wheel rut.
The distance moved is the sinkage Z, and the work done is given by
ER = 5.7cbz
where ER is the energy required to form the rut. Thus the rolling re-

Fig. 4.4. Wheel sinkage.


4.5 Bekker Theory 49

sistance R (assuming that it is due entirely to rut formation) is given


by
R = W/S.7cbd
For a driven wheel, the maximum (drawbar) pull can be obtained by
considering the thrust that can be developed given a value for the max-
imum shearing stress along the contact patch:
FD = H - R
Thus, for frictionless soils that have a maximum shear stress equal to
c, the drawbar pull is
FD = cbr sin e- W 2 /S.7cbd
where e is the angle of sinkage given in Fig. 4.4.

4.5 BEKKER THEORY

Fundamental to the Bekker approach to the theory of land traffic is


the relationship between the sinkage z and the normal pressure p. The
relationship developed by Bekker is a modification of an assumed linear
relationship between pressure and sinkage that is used in civil engi-
neering soil mechanics for small sinkages. Bekker (1960) gives the fol-
lowing equation:

p = (~ + k~ )zn
where ke is the cohesive modulus of sinkage, k~ the frictional modulus
of sinkage, and n an exponent reflecting the hyperbolic shape of the
load sinkage curve.
The values of ke, k~, and n can be determined for any given soil by
conducting load sinkage studies on two plates with different areas.
Log-log plots of pressure against sinkage will give straight-line rela-
tionships of slope n. Two equations for p at z = 1 enable values for
ke and k~ to be obtained.
The horizontal shear stress is given by a modification of the Cou-
lomb-Micklethwaite equation. For plastic soils, the following rela-
tionship is given by Bekker (1969):
,. = (c + p tan <1»(1 - e-j / k )
50 4. Traction

.J:::
C,
c
~ Plastic
(j) soil
(ij
Q)
.J:::
C/) Brittle
soil

- - - - - - <..
- J
Soil deformation

Fig. 4.5. Shear stress-deformation curves for plastic and brittle soils [from Bekker
(1969)].

where k is a slip coefficient and j the amount of soil deformation that


produces stress 1'. The Coulomb constants c and <\> can be determined
for a given soil by plotting maximum shear stress against normal pres-
sure to give the straight-line equation discussed in Section 4.4. The slip
coefficient (also termed the deformation constant) can be obtained from
stress-deformation curves that are obtained with a bevameter. * Figure
4.5 shows shear stress-deformation curves for plastic and brittle soils
(Bekker, 1969). If an annular shear ring is used, then the shear strength
is measured in terms of the shear torque, and the displacement is mea-
sured in terms of the angular deformation. These tests are usually re-
peated at various levels of normal pressures. Idealized shearing stress-
deformation graphs for different normal pressures are shown in Fig.
4.6. The yield points I, II, III, IV mark the end of the quasi-elastic
deformation and the beginning of plastic flow. This yield point, for a
given soil, will occur at the same deformation value and represents the
constant K. The idealized shear stress-deformation curves may not be
obtained in practice. On soft ground, and at high loads that are greater

* A bevameter is a device used to determine, in situ, the pressure-sinkage relationship


for a given supporting surface. It is usually mounted to a vehicle subframe, which acts
to provide force reaction during the penetration test.
4.5 Bekker Theory 51

Shear T
stress

Deformation j

Fig. 4.6. Idealized shearing stress-deformation curves.

than the bearing capacity of the soil, the shear stress may continue to
rise with deformation without giving any exact yield point. This situ-
ation requires correction of the previous equation for 'T. The extension
of the equation is not described here, but is fully discussed by Bekker
(1969) in his authoritative text dealing with terrain vehicle mechanics.
Slip i can be expressed in terms of the deformation} and the distance
x measured from the start of the ground contact area and some location
along the ground contact area:
} = ix
The thrust H at a particular slip can be found by using the above re-
lationship and the equation for horizontal shear stress. Integration over
the track length I gives

Hi = {hie (I + ~h) + Wtan ~[I + 0.64 (~) cot- ~J} 1

x (I + ~e-il/K
it
-~)
it
where Hi is the thrust at slip i.
The rolling resistance RJ is determined by considering the work done
in making a rut of length I and of depth z:

RJ = hi LZ p dz
52 4. Traction

Substituting for the radial pressure p gives


zn+ I )
R = ( ;+1 (ke + kq,b)

For a wheel, Bekker provides mathematical approximations for the


sinkage z and the rolling resistance R, such that
3W ] 2/(2n + I)
[
Zwheel = (3 - n)(ke + bkq,)YD

where D is the wheel diameter, and

= (3 - n)(2n+2)/(2n+ I)(n + l)(ke + bkq,)1I(2n+ I)

T
Rwheel

X [~ 2n +2)/(2n + I)

The drawbar pull FD can then be determined from


FD = Hi - Rwheel

A slightly modified version of the equation for the tractive force


developed by a tire was given by Bekker (1956) in terms of the slip i,
where
actual velocity
i = 1 - ------=----
theoretical velocity
and

H = (Ae + W tan <I>{ 1 - fz (1 - e- ilIk ) ]

where I is the length of the contact area (= 2[a(d - a)p/2), a the tire
deflection, d the tire diameter, and K the tangent modulus of defor-
mation from the ring shear test.

4.6 DIMENSIONAL ANALYSIS

In order to reduce the number of variables to be considered, a num-


ber of researchers have used dimensional-analysis techniques in de-
veloping prediction equations for the performance of wheeled vehicles.
4.6 Dimensional Analysis 53

Wismer and Luth (1974) identified nine pertinent variables relevant to


the prediction of traction: towed force (TF), pull (P), torque (Q), wheel
load (W), tire section width (b), overall tire diameter (d), tire rolling
radius (r), cone index (CI), and slip (0.
The Buckingham P theorem requires, by considering the number of
variables and the number of basic units, seven dimensionless ratios.
An appropriate set of dimensionless ratios was chosen by Wismer and
Luth as
TF P
W' W'
Three dependent-variable relationships were established by simili-
tude methods of experimental analysis:
1. Towed force (TF)

TF = 1.2W 004
W Clbd + .
2. Pull parallel to the ground (thrust, P)

~ = 0.75 (1 - e- O. 3 Cni) - (~~ + 0.04)


where Cn = ClbdlW.
3. Wheel torque (Q)

g = 0.75(1 - e- O. 3 Cni)
rW
The following assumptions are pertinent to the Wismer and Luth
equations: Axle torque was assumed to be zero for towed (i.e., non-
driven) wheels. Nominal tire pressure was assumed and is defined as
the pressure that produces tire deflections of approximately 20% of the
undeflected section height. The equations were developed for a width
to diameter ratio bid = 0.3.
Cone index, determined from cone penetrometer tests (ASAE, 1982),
was used as the measure of soil strength. Although this is a less accurate
method than the shear tests, it can be used quickly and easily in the
area in which vehicle traffic is to occur. In order to quantify the soil
strength adequately in terms of cone index, it is necessary to take a
number of individual measurements. The wheel numeric Cn is obtained
by obtaining the average cone index in the 0-150 mm layer.
54 4. Traction

4.7 TRACTIVE EFFICIENCY (1']D)

The tractive efficiency is the ratio between the output power obtained
from a wheel to the input power required to drive the wheel. It can be
expressed as follows:
1']0 = Fo V t (1 - i)IQw

where Fo is the pull force, V t the theoretical speed (= wr), w the angular
velocity of the wheel, r the rolling radius of the tire on a hard surface,
and i the slip.
For a two-wheel drive vehicle, the expression for tractive efficiency
is given by (Alcock, 1983)

T2 = [1 - (~~ + 0.04) + (~- 1)(~~ + 0.04) (1 - RSliP )]

x [0.75(1 - e( -O.3(C2)(R,h P )] - I

where T2 is the tractive efficiency of a two-wheel drive vehicle, C2 the


rear-wheel numeric, C 1 the front-wheel numeric, R slip the rear-wheel
slip (decimal), and R the ratio of the dynamic rear-axle weight to total
weight.
The effects of slip on the tractive efficiency of a two-wheel drive
vehicle are illustrated in Fig. 4.7 for Cn values ranging from 10 to 60.
A similar expression for the tractive efficiency of a four-wheel drive
vehicle (Alcock, 1983) is given as follows:

T4 = DBP/[ FAP + RAP ]


(1 - F s1ip ) (1 - R slip )

where T4 is the tractive efficiency, F s1ip the front-wheel slip (decimal),


FAP the front-axle pull, RAP the rear-axle pull, and DBP the four-
wheel drive drawbar pull.
The relationship between T4 and slip is shown in Fig. 4.8. For the
purposes of illustration, the same slip values were assumed for all the
driving wheels, and the rear-wheel numeric was set five units higher
than the front-wheel numeric, reflecting the compaction effect of the
front-axle driving wheels. It can be seen that the four-wheel drive ar-
rangement provides for higher levels of tractive efficiency, particularly
under poor soil conditions.
4.7 Tractive Efficiency (l1D) 55

Tractive
efficiency

0.8 ... -;..-1:.-::"":1:.: __


f!f'" ...... - -.. .:-::.....~ ......

.------...
.... .... tr ---~---.....-~-:.-:;:~
-
'-=:-;;...:>&,

0.6 ,
/ ....-- ~
----.. .......
,I ,,/ - .......... __ ...::::.,
<I ...... ... ...- .......::--.. .
,: ... -.... :::-........ ' .
1 --....
0.4 I

!
/
i - Cn =10
.--C n=20
0.2 J +--~ Cn=30
. -.. Cn=40
&-~Cn=50
_-.. Cn =60

o 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Wheelslip

Fig. 4.7. Tractive efficiency vs. wheelslip: two-wheel drive.

The expression given by Wismer (1982) for the tractive efficiency of


a single wheel is
=1- [(1.2/Cn ) + 0.04] (1 - i)
1]D 0.75(1 _ e-O.3Cni)
An equation for maximum pull was obtained by differentiating the
Wismer and Luth equation (Voorhees and Walker, 1977)

~ = 0.75(1 - e-O.3Cni) - (~~ + 0.04)

Thus,

:~ = 0.71 - (0.225C ni + 0.75)e-O.3Cni) - ~~


Voorhees and Walker suggested that this equation can be solved for
W (note that W is contained in Cn) by setting aPlaW = 0.075. This is
56 4. Traction

0.9.,-------.r------------------,

;,I,;!1~~
Tractive

~/"...-..--·--.. .~t
efficiency

0.8 //,1/
'T-::_ ~
:~~
t ,~ '-
/1, ,
~,t:
0.7 '"
1/:
,,:
./ :
0.6
Ii
f:
6:,,
, I

.............. C, = 15, C2=20


,, . - ... C, =20, C2=25
0.5 ,
I ~-~ C, =25, C2=30

,: .. - . C, =30, C2=35
e--e C, =35, C2=40
f .... --... C, =40, C2=45
0.4 •

o 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4


Wheelslip

Fig. 4.8. Tractive efficiency vs. wheelslip: four-wheel drive.

regarded as a point on the pull-wheel load curve that is practically


equal to zero. The value of W obtained is the weight required to obtain
maximum pull. It can then be used in the Wismer and Luth P/W equa-
tion to determine the maximum pull.

4.8 MOBILITY NUMBER

Gee-Clough et al. (1978) used a dimensionless parameter termed


"mobility number" in developing empirically based traction prediction
equations similar to those of Wismer and Luth. The mobility number
is, in fact, an extended form of the wheel numeric en.
Mobility number M is defined as

= Clbd (~) 112 ( 1 )


M W h 1 + bl2d
where CI is the soil cone index, b the tire width, d the tire diameter,
4.8 Mobility Number 57

W the tire load, h the tire section height, and athe tire deflection under
load.
Wismer and Luth plotted the dimensionless term PIW against slip,
whereas Gee-Clough et al. plotted the same term against mobility num-
ber. It should be noted that the dimensionless quantity PIW is referred
to as the "coefficient of traction" in Europe, whereas the American
Society of Agricultural Engineers (ASAE) has adopted the expression
"pull ratio," symbol /-L, for the same term.
The empirical relationship for the pull ratio (coefficient of traction)
C T established by Gee-Clough et al. (1978) is of the form
C T = (CT)max(1 - e- ki )

where i is the slip, k the rate constant, and (CT)max the maximum
coefficient of traction, and the maximum coefficient of traction was
found to be
(CT)max = 0.796 - 0.921M
For the rate constant,
k(CT)max = 4.838 + 0.061M
Wismer and Luth used the dimensionless ratio TFIW in predicting
the required towed force (TF). This ratio is also referred to as the
"coefficient of rolling resistance" (CRR ). Gee Clough et al. (1978) gave
the following relationship between the coefficient of rolling resistance
and mobility number:
C RR = 0.049 + 0.2871M
There are obvious similarities between the equations developed by
Gee-Clough et al. and those developed by Wismer and Luth. The trac-
tive efficiency 'YjD relationship to slip can be established as follows:
CT (1 - i)
'YjD =
CT + C RR
It must be pointed out that the relationships developed by Gee-
Clough et al. are for field conditions such as stubble, ploughed, and
cultivated ground. They do not apply to hard, dry surfaces, which give
higher tractive results, or to very slippery conditions, which give poor
tractive performances.
Traction and tractive performance is a fundamentally important as-
58 4. Traction

pect of agricultural tractor performance. This chapter has reviewed


methodologies for predicting tractive capabilities based on physical and
empirical relationships.

REFERENCES
Alcock, R. (1983). Battery powered vehicles for field work. Trans. ASAE 26(1), 10-13.
ASAE (1982). Soil cone penetrometer. ASAE Standard S.313. Am. Soc. Agric. Engr.
Bekker, M. G. (1956). "Theory of Land Locomotion." Univ. of Michigan Press, Ann
Arbor.
Bekker, M. G. (1960). "Off-the-Road Locomotion." Univ. of Michigan Press, Ann
Arbor.
Bekker, M. G. (1969). "Introduction to Terrain Vehicle Systems." Univ. of Michigan
Press, Ann Arbor.
Gee-Clough, D., McAllister, M., Pearson, G., and Evernden, D. W. (1978). The empirical
prediction of tractor-implement field performance. 1. Terramech. 15(2), 81-94.
Micklethwaite, E. W. E. (1944). Soil mechanics in relation to fighting vehicles. Military
College of Science, Chertsey, England.
Reece, A. R. (1966). Principles of soil-vehicle mechanics. Proc. Inst. Mech. Engr. 180,
Part 2A(2), 45-66.
Uffelmann, F. L. (1961). The performance ofrigid cylindrical wheels on clay soil. Proc.
1st Int. Conf. Mechanics of Soil-Vehicle Systems, Turin.
Voorhees, M. L., and Walker, P. N. (1977). Tractionability as a function of soil moisture.
Trans. ASAE 20(5),806-809.
Wismer, R. D. (1982). Soil dynamics: A review of theory and applications. SAE paper
no. 820656. SAE, Warrendale, PA.
Wismer, R. D., and Luth, H. J. (1974). Off-road traction prediction for wheeled vehicles.
Trans. ASAE 17(1),8-10, 14.
5
Tires and Ballasting
for Agricultural Tractors

5.1 TIRE DEVELOPMENT

The steel-wheeled tractor was replaced by vehicles fitted with pneu-


matic tires in the 1920s and 1930s. Harvey Firestone (1868-1938) was
instrumental in developing reinforced, "low-pressure" tires that were
suitable for agricultural field work. World War II resulted in shortages
of natural rubber previously used for tire production. This accelerated
the research and development of synthetic rubbers such as styrene and
latex. The synthetic rubbers produced had better wear-resistant prop-
erties than the original rubber tires. Synthetic rubbers that are com-
monly used today include polyurethane, neoprene, polybutandiene,
and butyl. Natural rubber may still be used for some specialized ap-
plications.

5.2 TIRE FUNCTION

The agricultural tire must perform the following functions:


1. support the vehicle and associated loads at some low level of
ground pressure,
59
60 5. Tires and Ballasting for Agricultural Tractors

2. absorb shock loads and cushion the vehicle against minor surface
irregularities,
3. provide traction (and braking),
4. provide for steering and directional stability,
5. resist the abrasive action of the various surfaces on which it is
expected to operate.

5.3 TIRE TERMINOLOGY

The prefix R has been adopted by ASAE (1983) as the code for rear,
or drive, tires. The prefixes F and I have been adopted for steering
tires (or front tires) and for implement tires, respectively. The "reg-
ular" agricultural drive tire is coded R-l. There are three other cate-
gories of drive tire: cane and rice (R-2), industrial and sand (R-3), and
industrial tractors (R-4). Front or steering tires are coded F-l for single-
rib tires, F-2 for "regular" agricultural tires, and F-3 for industrial
multiple-rib tires. Implement tires are coded 1-1 for ribbed, 1-2 for mod-
erate traction, 1-3 for traction, and 1-6 for smooth implement tires.
Tire sizing is based on the tire section width and the rim diameter.
Thus a 18.4-26 drive tires has a section width of 18.4 in. and a rim
diameter of 26 in. Early tire designs had tire section widths equal to
the tire section height. The aspect ratio of these tires, as defined by
h/b, where h is the tire section height and b the tire section width, was
thus equal to 1.0. The length of the contact patch is given by
1= 0.31d (5-1)
where I is the length of the contact patch and d the diameter of the
unloaded tire. The contact patch can be regarded as elliptical in shape,
and the cross-sectional area of the patch is given by
A = 'ITlb (5-2)
where b is the width of the contact patch.
Today's tires have wider section widths, and aspect ratios of 0.85
or even 0.75 are found (Inns and Kilgour, 1978).
The ply rating of a tire is used to indicate its load-carrying ability.
The ply rating was once used to specify the number of plies built into
the construction of the tire. This is no longer the case, but the term
has been retained as an index of the tire strength. There are three types
of casing construction: cross ply, radial ply, and belted bias (Anslow
5.4 Ballas ling 61

and Warrilow, 1970). In the case of cross-ply construction, the cords


are arranged at an angle of approximately 40° to the circumferential
centerline of the tire. This is referred to as the crown angle (Wong,
1978). There are normally two or more layers or cords, or plies, each
layer being set in the opposite bias. The majority of agricultural tires
are of this construction. The advantage of this construction for off-road
vehicles is that the sidewalls have reinforcement, thus providing some
protection against their impact damage. Ply tires are constructed using
individual cords running from bead to bead in a line perpendicular to
the circumferential centerline of the tire. Thus the crown angle is 90°.
There are also a number of belts consisting of several cords, which are
fitted on top of the radial plies. The cords within these belts form a
crown angle of 20°. The belts are provided to brace the tread and pre-
vent buckling around the periphery of the tire. The radial-ply tire pro-
vides, relative to the cross ply, uniform ground pressure over the con-
tact area. The length of the contact patch relative to an equivalent
cross-ply tire is longer, and thus the flotation characteristics of the
radial tire are better. The radial tire is also likely to develop higher
levels of drawbar pull than the equivalent cross-ply tire, particularly
on light, sandy loam soils. The cross ply may show some improvement
in performance relative to the radial ply, in terms of pull-slip char-
acteristics, on wet, heavy clay soils (Anslow and Warrilow, 1970). The
belted bias is designed as a combination of the radial- and cross-ply
tire designs, although the principal plies are arranged in cross-ply con-
figeration, set at a low crown angle.

5.4 BALLASTING

The tractive performance of a tractor can be improved, particularly


on sandy loam soils, by ballasting, which adds weight to the drive
wheels. It is important, however, that the tire manufacturer's rec-
ommended load, at normal inflation pressures, not be exceeded. The
addition of ballast to a tractor is usually accomplished by adding wheel
weights, adding front weights, and filling the tires with water. Adding
too much ballast will result in excessive power loss due to increased
rolling resistance, but insufficient ballast will cause power loss because
ofthe increased wheelslip. The driving weight required over each trac-
tion tire can be determined from the traction prediction equations of
Wismer and Luth (Chapter 4). It is important to match the tractor
62 5. Tires and Ballasting for Agricultural Tractors

power, weight, speed, and draft force. For a two-wheel drive tractor,
Gee-Clough et al. (1982) cite Reece (1970) as given the relationship
W*/P = 1.17/V (5-3)
and for four-wheel drive tractors
W*/P = 0.82/V (5-4)
where W* is the total tractor weight in kilo Newtons (kN), P the engine
power (kW), and V the forward speed (m sec-I).
Dwyer (1978) gave a relationship of the form
W/p* = 1.79/V (5-5)
where W is the dynamic weight on the drive wheels (kN) and P* the
total axle power (kW). This expression was obtained using average
values from field tests. The equation is intended to give the weight on
the drive tires per unit of axle power to ensure operation at maximum
efficiency. Dwyer suggested operating at a slip of 0.1 (10%) and a coef-
ficient of traction of 0.4.
If PTO power is to be used instead of axle power, then the losses
between the axle and the PTO must be taken into account. Since PTO
power levels are readily available from test reports, there is some prac-
tice value in relating dynamic weight to PTO power output. For a trac-
tive efficiency of 0.7 and a coefficient of traction of 0.4, the PTO power,
weight on the driving wheels, and the vehicle speed can be related as
follows (Bloome et al. 1983):
0.7(PTO power) = O.4(weight on drive wheels)(vehicle speed) (5-6)
This gives the relationship WV = 1.79 P* of Eq. (5-5). Other values
for coefficient of traction and tractive efficiency can be used to give
different relationships. If a 4% loss is assumed between the PTO power
and the axle power, then the relationship given in Eq. (5-5) can be
written as
km hr- 1 kN (PTO kW) - 1 = 6.20 (5-7)
Bloome et al. (1983) suggest that this equation is applicable to deter-
mining the optimum ballast for power-limiting conditions. The under-
ballasted tractor is traction limited, and at a wheelslip of 20% and good
soil conditions, a coefficient of traction of 0.5 is typical.
Assuming a tractive efficiency of 0.7 gives an expression
W/p* = 1.4/V (5-8)
5.4 Ballasting 63

or
km hr- I kN (PTO kW) - I = 4.84 (5-9)
Bloome et al. (1983) discussed ballasting recommendations for two-
and four-wheel drive tractors. Dwyer (1978) and a tractor manufacturer
recommend the same mass-to-power relationships for two- and four-
wheel drive tractors operating at the same speed. There is also a case
for having a slightly greater mass-to-power ratio for two-wheel drive
tractors, since some axle force must be maintained on the unpowered
front wheels to provide steering control. Alternatively, it can also be
argued that four-wheel drive tractors can use slightly greater mass-to-
power ratios since there is no rolling resistance loss associated with
unpowered wheels. Bloome et al. (1983) conclude that ballasting rec-
ommendations should be the same for two- and four-wheel drive trac-
tors operating at the same speed. The ballast recommendations based
on Dwyer (1978) are for utilization of full engine power. Tractors op-
erating at less than full rated power will require some reduction in
tractor ballast. This is perhaps best achieved for these part-load con-
ditions by assuming appropriate values for tractive efficiency, coeffi-
cient of traction, and forward speed, and subsequently modifying Eqs.
(5-7) and (5-9) in accordance with Eq. (5-6).
Gee-Clough et al. (1982) developed ballast recommendations based
on the traction prediction equations developed by Gee-Clough et al.
(1978). These equations were discussed in Chapter 4 and predict coef-
ficient of traction and coefficient of rolling resistance, based on mobility
number M. Using these equations, Gee-Clough et al. (1982) developed
an approach for estimating the theoretical loss in tractive efficiency
resulting from having drive tire loads greater or smaller than the op-
timum value.
The power transmitted by the drive wheels, PI (kN), is given by
(5-10)

where W is the dynamic load on the drive wheels (kN), CT the coef-
ficient of traction, and V the forward speed (m sec - I). The maximum
power that the wheels are able to transmit, P 2 (kW), at any value of
slip is given by
(5-11)

where T) is the tractive efficiency and P* the total axle power (kW).
64 5. Tires and Ballasting for Agricultural Tractors

If PI = P 2 , then the power transmitted will be equal to the maximum


possible at that value of slip, and
WCT V = 1]P* (5-12)
or
W=..2l...! (5-13)
P* CT V
In order for operation at maximum power to be achieved, the slip at
which PI = P 2 has to be the slip at maximum efficiency. Gee-Clough
et al. (1982) showed that for the range of mobility numbers normally
encountered for soil conditions (8-30), the slip at which maximum ef-
ficiency occurs is 10%. If this represents the optimum condition, then

P~Pt = (~)oPti (5-14)


combining Eqs. (5-13) and (5-14) gives
(W/p*) (CT)opt 1]
(W/P) opt = ----c;-
1]opt
(5-15)

For preset values of wheelslip and mobility number, values of W/P*,


CT , and 1] can be found. These can be compared to the optimum values
and a curve of W/Wopt against 1]/1]opt drawn. Gee-Clough et al. (1982)
produced the diagram of this relationship, which is reproduced in Fig.
5.1. It is apparent that there is a desirable range for W/Wopt, but that
it can be allowed to deviate from W/Wopt = 1, particularly at the in-
creased values for mobility number.
If the power actually transmitted by the wheels, PI, is always less
than the maximum power able to be transmitted, P 2 , at all values of
slip, then the tractor will never be power limited. The conditions under
which this would occur are calculated from Eq. (5-10) as
PI = WCTV = WCT V O(1 - i) (5-16)
where Vo is the theoretical (no-load) forward speed. From Eq. (5-11),

P2 = 1]P* = CT (1 - i) P* (5-17)
CT + CRR
and

(5-18)
5.4 Ballasting 65

1.0
/
/
""
I " "-
I
"" ,M=3
0.9
I
I
,,
I
,,
'l/'loPt I
,
,
I
0.8 I

I
I
0.7 I
I
I
I
0.6

0 0.4 0.8 1.2 1.6 2.0


W/W opt

Fig. 5.1. Ratio 1l!1lopt vs. W!WoPt for mobility number M = 3 and M = 30 [from Gee-
Clough et al. (1982)].

The expression WVo/P* is a dimensionless number, which is termed


the ballast number by Gee-Clough et al. (1982). The tractor drive tires
will never be able to deliver the power available if the ballast number
is smaller than that required to give PI = P 2 • This is the power-limited
condition. Gee-Clough et al_ (1982) showed that if the maximum slip
is assumed to be 20%, then the minimum value for the ballast number
to ensure the tires deliver the available power is 1.58. Thus

W = 1.58 (1 - 0.2) 1.26


(5-19)
P* V V
This can be rewritten as
km hr- I kN (PTO kW)-1 = 4.36 (5-20)
Equation (5-20) provides for a ballast level 70% of that given for the
optimum ballast in Eq. (5-7). Therefore, the minimum ballast recom-
mendation should be at 70% of that for the optimum level. Experiments
by Gee-Clough et al. showed no significant reduction in power output
up to a maximum ballasting of 70% of the optimum. Thus an acceptable
recommendation for the ballast is that the dynamic tire load should be
66 5. Tires and Ballasting for Agricultural Tractors

70-140% of the optimum value to avoid major losses in the output


power at the drawbar. The maximum ballast can be written as
km hr- I kN (PTO kW)-1 = 8.68 (5-21)
To summarize:
1. The coefficient of traction at maximum efficiency can be as-
sumed to be 0.38-0.4 for most soil conditions encountered in agricul-
ture.
2. This level of coefficient of traction will give maximum efficiency
at a slip of 10%.
3. The load on the drive tires should be based on axle power and
working speed in accordance with Eq. (5-7). That is,
km hr- I kN (PTO kW)-1 = 6.20
The load on the tires is the dynamic tire load. It includes the static
load and the weight transfer effect.
4. The minimum ballast for the traction-limited condition is given
by a dynamic load that is 70% of the optimum. This is given by Eq.
(5-20) as km hr- I kN (PTO kW) - I = 4.36.
5. The maximum ballast for the power-limited condition is given
by Eq. (5-21) as km hr- I kN (PTO kW)-1 = 8.68.
A weight transfer analysis specifically of four-wheel drive tractors was
presented by Peters (1983). The equations developed were used to ex-
press the vehicle performance in terms of the rear axle. Figure 5.2
shows the free-body diagram used by Peters in analyzing the forces on
a four-wheel drive tractor.
Summing moments about the rear axle gives

(5-22)

where RI is the front dynamic load, W the vehicle weight, Q3 the total
axle torque ( = QI + Qz), and Po the resultant drawbar pull.
The front static weight is given by the term W(XdX z ) and the load
transfer component is given by
Q3 - (X4P o)IXz
The ballast requirement can be determined from the coefficient of trac-
tion. Equations for coefficient of traction (which is also referred to as
5.4 Ballasting 67

a,

Veh icle
net
tract ion

Fig. 5.2. Forces acting on a four-wheel drive tractor [from Peters (1983)].

the "dynamic ratio" and "pull ratio") were given in Chapter 4. The
general expression for coefficient of traction is (Leviticus and Reyes,
1983)

and
(5-23)
where P is the drawbar pull, W the dynamic weight on the tire, b the
tire section width, d the tire diameter, i the slip, (CT)max the constant
equivalent to the maximum coefficient of traction, and K a constant
related to the tire resilience (kN/m 2 ). For a tire moving on soil, the K
factor is taken to be equivalent to the cone index.
Leviticus and Reyes (1983) developed an expression termed the
"tractive quotient" (TQ) from Eq. (5-23):

TQ = A - PIW
A
and thus
TQ = e-K(bdlW)i (5-24)
This expression was used to evaluate the tractive response for dif-
68 5. Tires and Ballasting for Agricultural Tractors

ferent tire-loading factors. Tire loading is represented by the expression


bd/W. Values for the constants A and K were obtained from the analysis
of Nebraska tractor tests. The results were therefore only applicable
to performance on concrete. As the tire-loading factor was increased,
the coefficient of traction was found to increase at the same value of
slip. However, the rate of increase of coefficient of traction was re-
duced at higher tire-loading factors.

5.5 TIRE SELECTION

A rational approach to tire selection based on coefficient of traction


determined from mobility number was proposed by Gee-Clough (1980).
This approach is based on the empirical models of Wismer and Luth
(Chapter 4) and of Gee-Clough et al. (1982). [The empirical equations
used to predict coefficient of traction CT , maximum coefficient of trac-
tion (CT)max, and coefficient of rolling resistance C RR are discussed in
Chapter 4, Section 4.8.] The dimensionless term mobility number (M)
is used in the empirical relationships. It is defined as

= Cbd (~) 1/2 ( 1 ') (5-25)


M W h 1 + b/2d
where C is the soil cone index value, b the tire width, d the tire di-
ameter, h the tire section height, 0 the tire deflection under load, and
W is the dynamic load on the tire.
The tire deflection 0 is normally measured statically on a hard sur-
face, and a typical value of o/h is 0.2 at the manufacturer's recom-
mended load and inflation pressure. Gee-Clough took CI readings over
several years and established values of 200, 700, and 1500 kPa to rep-
resent bad, average, and good field conditions, respectively. In order
to accommodate the effects of various operating parameters on the
prediction of tractive performance, correction factors for (CT)max, K,
and CRR were provided by Gee-Clough. These are reproduced in Table
5.1.
The correction factors are meant to be applied to a discrete situation
and should not be combined. It is interesting to note that no correction
factor is included for the tire aspect ratio. Change in aspect ratio will
affect tractive performance; however, Gee-Clough noted that the range
of aspect ratios commercially available is small and found no significant
5.5 Tire Selection 69

Table 5.1. Correction Factors for Tractive Performance Parameters U

Correction factor
Traction
Effect conditions (CT)max K CRR

Radial ply compared to cross ply Bad 0.95 1.38 1.00


Average 0.95 1.38 1.00
Good 0.95 1.38 1.00
High-lugged tires Bad 1.10 0.92 1.03
(i.e., 75 vs. 35 mm) Average 1.10 0.92 1.21
Good 1.10 0.92 1.32
Forward-speed increase Bad 1.01 1.00 0.98
(3.2 to 6.4 km/hr) Average 1.01 1.00 0.98
Good 1.01 1.00 0.98
Running in furrow bottom Bad
(instead of field surface) Average 1.25 1.00 1.00
Good 1.23 0.69 1.80

a From Gee-Clough (1980).

difference between a tire with an aspect ratio of 0.69 (advertised as a


low aspect ratio) and a tire with an aspect ratio of 0.75.
In the following example, traction parameters are calculated for an
axle carrying 14 kN, or 7 kN per tire. The traction parameters can be
calculated as follows:
For a 9.5 x 24 agricultural drive tire with a load of 7 kN applied,
the pertinent dimensions are
tire breadth b = 0.24 m
tire diameter d = 1.04 m
tire aspect ratio = 0.2
For bad conditions, the cone index C = 200 kPa.
The mobility number M is given from Eq. (5-25) as
. M = 200 x 0.24 x 1.04 ~ ~O2 ( 1 ) 2 85
.. 7 V V.L. 1 + 0.2412.08 .
The maximum coefficient of traction is given by
0.92 0.92
(CT)max = 0.796 - M = 0.796 - 2.85 0.47

The coefficient of rolling resistance is


0.287
CRR = 0.049 + 0.287 = 0.049 0.15
M + 2.85
70 5. Tires and Ballasting for Agricultural Tractors

Table 5.2. Predicted Tractive Performance Parameters for a 9.5 x


24 Agricultural Drive Tire

Tractive
conditions (CT)max k(CT)max C RR 1lmax

Bad 0.47 5.01 0.15 0.61


Average 0.70 5.43 0.08 0.74
Good 0.75 6.10 0.06 0.79

The slope of the coefficient of traction versus slip curve at the origin
is
K(CT)max = 4.838 + 0.061M = 4.838 + 0.061(2.85) = 5.01
The slip at maximum efficiency was shown by Gee-Clough (1980) to
vary little with mobility number M over the range of mobility numbers
encountered for agricultural soils. It was therefore suggested that a slip
value of 0.10 would represent a useful average figure for slip at max-
imum efficiency.
The tractive efficiency T) was given in Chapter 4 as

where CT is the coefficient of traction and i the slip.


The coefficient of traction CT is given by

where K = k(CT)max/(CT)max'
For the example given, K = 5.01/0.47 = 10.66. Thus,
0.308(1 - 0.1)
CT = 0.47(1 - = 0.308,
e-IO.66XO.l)
T)max = 0.308 + 0.15 = 0.61
The predicted tractive performance parameters for bad, average, and
good traction conditions are given in Table 5.2.
Tire selection for drive tires can be made on the basis of tires that
meet the load specifications and provide the highest tractive efficiency
under the ground conditions expected. This necessitates repeating the
5.5 Tire Selection 71

previous calculations for each tire option. If the tires are towed, then
the selection is on the basis of that tire size providing the minimum
coefficient of rolling resistance CRR • Tire options available will, how-
ever, be limited by the range of tire sizes commercially available, and
by design limitations imposed by the machine size, etc. Single and dual
tires can be compared by calculating the mobility numbers for each
situation (Gee-Clough, 1980). If the duals are assumed to act indepen-
dently of each other, to have width b, and to carry a load W12, with
diameter d, the mobility number is given by

MI =
Cbd
WI2
(5)h 112
1
I
+ (bl2d)
For the same tire width, the single tire will have a width 2b, diameter
d, and carry load W. The mobility number is given by

M2 = C(2b )d (~) 112 1


W h 1 + (2bI2d)
A comparison gives

MI = 1 + bid> 1
M2 1 + bl2d
The ratio is greater than 1, and thus duals will perform better than
singles of the same overall dimension.
This approach can also be used to compare large single tires with
smaller duals.
For duals of diameter d l , width b l , and load W12, the mobility number
is

MI -
_ Cbldl (~)1/2
WI2 h I
___
+
_ (btl2d l )

For single tires of diameter d 2 , width b 2 , and carrying load W, the


mobility number is
= Cb 2 d 2 (~) 112 _ _ __
W h I + (b 2 /2d 2 )
A comparison gives
72 5. Tires and Ballasting for Agricultural Tractors

If
2b 1 d 1
bzdz > (1 + b 1/2dd(1 + bz/2dz)

then MdM2 will be greater than 1. In this case, the duals will have
better tractive performance than the single tire.

REFERENCES
Anslow, B. J., and Warrilow, B. J. (1970). Tractor tires. Proc. Inst. Mech. Engr. 184,
Part 3Q, 170-182.
ASAE (1983). Tire selection table for agricultural machines of future design. ASAE
Standard S.220.4. Am. Soc. Agric. Eng. Yearbook, St. Joseph, ML
Bloome, P. D., Summers, J. D., Khalilian, A., and Batchelder, D. G. (1983). Ballasting
recommendations for two wheel and four wheel drive tractors. ASAE Paper No. 83-
1067. Am. Soc. Agric. Engr., St. Joseph, ML
Dwyer, M. J. (1978). Some aspects of tyre design and their effect on agricultural tractor
performance. Proc. Inst. Mech. Engr. Conf. Off-Highway Vehicles, Tractors and
Equipment. Institute of Mechanical Engineers, London.
Gee-Clough, D. (1980). Section of tire sizes for agricultural vehicles. J. Agric. Engr.
Res. 25,261-278.
Gee-Clough, D. McAllister, M., Pearson, G., and Evernden, D. W. (1978). The empirical
prediction of tractor-implement field performance. J. Terramech. 15(2), 81-94.
Gee-Clough, D., Pearson, G., and McAllister, M. (1982). Ballasting wheeled tractors to
achieve maximum power output in frictional-cohesive soils. J. Agric. Engr. Res. 27,
1-19.
Inns, F. M., and Kilgour, J. (1978). "Agricultural Tyres." Dunlop Ltd., London.
Leviticus, L. I., and Reyes, J. F. (1983). Traction on concrete. I, dynamic ratio and
tractive quotient. ASAE Paper No. 83-1558. Am. Soc. Agric. Engr., St. Joseph, ML
Peters, B. (1983). Weight transfer analysis on four-wheel drive tractors. ASAE Paper
No. 83-1068, Am. Soc. Agric. Engr., St. Joseph, ML
Reece, A. R. (1970). The shape of the farm tractor. Proc. Ins!. Mech. Engr. 184, Part
3Q,125-131.
Wong, J. Y. (1978). "Theory of Ground Vehicles." Wiley, New York.
6
Machinery Management

Machinery management is recognized as being important in reducing


the cost of crop production. For example, Ozkan and Edwards (1983)
state that' 'in Iowa, owning and operating agricultural machinery com-
prises 31 and 43 percent, respectively, of the total non-land costs of
corn and soybean production" (p. 2). Von Bargen and Cunney (1974)
suggested that over one-third of crop production costs can be attributed
to machinery operation. There are, however, obvious difficulties in-
herent in any proposal to determine machinery requirements. The se-
lection of farm machinery, and decisions relating to replacement will
be based on some economic criteria, and yet the absolute cost of a
machine can only be determined after its replacement. This means that
although we cannot be specific in presenting machinery selection cri-
teria, reasonable estimations can be made.

6.1 FIXED COSTS

The costs of owning farm machinery can be divided between fixed


and operating costs. Fixed costs generally are those not affected by
the amount of use of the machine, for example, interest, depreciation,
shelter (housing), or insurance. Ownership costs can be considered as
73
74 6. Machinery Management

those that vary with the amount of use of the machine, and would
include items such as fuel, tires, labor, lubrication, and repairs.
A number of methods for calculating the costs of owning farm ma-
chinery are available. Of importance to this text is not the method of
calculating, but how this information is then utilized to assist in farm
machinery decisions. Hunt (1973) uses a fixed-cost percentages factor
in determining the annual cost of operating a machine. He lists the
value of the fixed-cost percentage as a function of service life, where
service life is defined as the length of time (in years) until the machine
is only of value for salvage. The fixed-cost percentage is then applied
(as a product) to the purchase price of the machine, thus providing a
relatively simple method of determining fixed costs.
The fixed-cost percentage (FC%) can be determined from (ASAE,
1983a)

FC% = [1-
lOOP - n -S + (1- 2+ -s i) + (s - i) ]

where P is the purchase price of the machine, s the percentage salvage


value, i the percentage interest rate, and n the machine life in years.
A more sophisticated approach can take into account interest rates,
tax and inflation, machine life, and resale or trade-in value. The mean,
after-tax cost of machinery ownership FC AT can be expressed by the
following equation (Bartholomew, 1976):
FC AT = [(P - A - D - R)I] + H
where P is the current price of the equipment, or an equivalent ma-
chine, A the effect of an investment credit, D the cumulative effect of
depreciation allowance, R the effect of reselling or trading-in the equip-
ment, I an annuity factor to give a fixed annual charge over the life of
the equipment, which is equivalent to its net present value, and H an
allowance for housing and insurance of the machine. Items P, A, D,
and R are reduced to a net present value.
The effect of the investment credit, A, calculated at rate a (which is
currently 0.1 for agricultural machinery purchases), can be calculated
as benefit received, discounted for interest and inflation:
Pat
A = (1 + k)(1 + g)
where t is the average tax rate, k = i(1 - t) the after tax cost of
borrowing, and i the effective annual interest rate.
6.1 Fixed Costs 75

Table 6.1. Resale Price Constants

f K

Tractors 0.68 0.920


Combines, cotton pickers, 0.64 0.885
self-propelled windrowers
Balers, forage harvesters, 0.56 0.885
self-propelled sprayers
Other field machines 0.60 0.885
Equipment with zero resale value

The depreciation allowances D are summed over each year to a final


year m, and discounted for interest and inflation:

D = ~
n~ I
t?n
(l - k) (l - gy
where Dn = dS(n-l) is the depreciation allowance in year n, with d
the allowable diminishing-value depreciation rate and S(n-I) the de-
preciation value at the end of the previous year.
The effect of reselling or trading-in the machine in year m is given
by
R = Rm(l + g)m + t{Sm - [Rm(l + g)m]}
(l + k)m(l + g)m (l + k)m(l + g)m
where Sm is the written-down value and Rm the resale price in year m.
The estimated resale price can be obtained from

Values for constants f and K are given in Table 6.1 (ASAE, 1983a).
The annuity factor I for converting a net present value, or cash outlay,
to an equivalent annual charge is given by
I = k(l + k)m
(l + k)m - 1

The allowance H for insurance and housing is given by


H = (1 - t)(h + v)P
where v is the insurance rate and h the housing rate, with both ex-
76 6. Machinery Management

pressed as a proportion of the new price of the machine. By way of


example, h + v is commonly assumed to be in the range 0.02-0.03.

6.2 OPERATING COSTS

Operating costs OC can be expressed by the following relationship:

OC = (1 _ t) ~ Fn + Ln + Mn + Rn
n-=1 (1 + kt I
where
I = k(1 + k)m
(1 + k)m - 1

is the annuity factor, Rn the annual repair cost, Mn the annual main-
tenance cost, Ln the annual labor cost, Fn the annual fuel cost, t the
average, or marginal, tax rate, k the after-tax cost of capital, and k =
i(1 - t) with i the interest rate and m the number of periods.

6.2.1 Fuel Consumption


The fuel consumption FC for a particular operation can be estimated
from the following equation (ASAE, 1983b) (assuming that power is
supplied by a diesel engine tractor):
FC = 2.64X + 3.91 - 0.2Y738X + 173 liters/Kwh
where X is the ratio of equivalent PTO power required to the maximum
available from the PTO. The average annual fuel consumption FC can
be estimated by (for a diesel engine tractor)
FC = 0.223PTO max liters/hr

6.2.2 Maintenance Costs


Maintenance costs are likely to be extremely variable and therefore
difficult to specify in terms of an estimation. Maintenance, as distinct
from repair, is an action taken to prevent the equipment from reaching
a failed state. The costs associated with maintenance are thus influ-
enced by the frequency with which maintenance procedures are carried
6.2 Operating Costs 77

out. Maintenance procedures are usually invoked, not to restore the


equipment to its original condition, but to prevent as far as possible,
untimely breakdowns. The restoration of the equipment to its "as new"
condition can be considered as an overhaul. This implies that, if main-
tenance procedures only are followed, the equipment will gradually
deteriorate over time (presumably to the point at which the equipment
requires overhaul or replacement). This approach to defining the ef-
fective life of the machine is built into a replacement model that is
discussed in Section 6.5. The effect of regular maintenance should be
to increase the mean time to failure of a component (or the equipment).
The ASAE Yearbook (l983c) provides a table for the estimation of
machinery repair and maintenance costs based on the accumulated use
of the machine. The table gives values for the accumulated repair and
maintenance costs as a proportion of the list price of the machine.
The question of maintenance routine is linked to the problem of
determining the frequency with which maintenance routines should be
conducted so as to minimize the downtime. A mathematical model that
predicts the maintenance frequency can be constructed if the following
assumptions are made (Jardine, 1973):
1. Equipment failures occur according to a negative exponential
distribution, where A is the mean arrival rate of failures, and 1IA the
mean time to failure. The negative exponential distribution arises where
failure of anyone of a number of components causes the equipment
to stop. It is also a characteristic of equipment subject to failure due
to random causes.
2. Repair times occur according to a negative exponential distri-
bution with a mean time of 1IA.
3. The maintenance policy is to perform n inspections per unit time.
Inspection times are negative exponentially distributed with a mean
time of lIi.
4. The breakdown rate A is a function of n (the frequency of main-
tenance).
For purposes of illustration, it is assumed here that the breakdown
rate varies inversely with the frequency of maintenance. That is,
A(n) = kin

where the constant k can be regarded as the arrival rate of breakdowns


78 6. Machinery Management

per unit time when one maintenance check is carried out per unit time.
Differentiating,

d (Mn)) = A'(n) = -k
dn n2
The total downtime D(n) per unit of time will be a function of the
maintenance frequency, and so
D(n) = downtime due to repairs per unit time
+ downtime due to maintenance per unit time
The proportion of time the equipment spends being repaired is given
by Mn)/f.L and the downtime due to maintenance per unit time is n/i.
Thus

D(n) = Mn) +~
f.L I

The minimum point is obtained by setting the derivative of D(n) equal


to zero:
-k
D'(n) = -2
1
+ -:- = 0, or n = (~) 112
n f.L I

Let the average number of breakdowns per week when one mainte-
nance routine is conducted be 2. Then k = 2. Let the mean time to
failure be 1/f.L = 16 hr or 0.095 weeks. Maintenance/inspection times
have a mean of lIi = 4 hr or 0.024 weeks. Then

n (ki) = (2 x10.441. 7) =2.8


=; 112 1/2
per week

This result could be interpreted as approximately three maintenance


routines per week.

6.3 RELIABILITY

Reliability is the probability that the machine or machine system will


operate as expected for a specified period of time. In the consideration
of maintenance we are interested in the probability of a failure occur-
ring before some specified time t. Reliability, however, is concerned
with the probability that the equipment will survive to at least time t.
6.3 Reliability 79

The probability of failure before time t is termed the cumulative dis-


tribution function F(t), where

F(t) = f~x f(t) dt.

As t ~ 00, F(t) ~ 1. The reliability function R(t) is the complement of


the cumulative distribution function:

R(t) = 1- F(t) = f'" f(t) dt

As t ~ 00, R(t) ~ O.
The failure rate is an important statistical characteristic used in re-
liability studies. It is the ratio of the number of items that fail in an
interval of time to the number of items that constituted the original
population. The probability of an item failing in time interval ot, given
that it has survived to time t, is represented by r(t)dt.
The failure rate in interval ot is

( )
<:: _ F(t + ot) - F( t)
r t ut - 1 _ F(t)

The instantaneous failure rate is given by considering ot ~ 0, giving


(Pitt, 1980)
f(t)
ret) = 1- F(t)

where f(t) is the probability density function.


Figure 6.1 shows the form that is often assumed for the failure rate
of complex equipment. The shape of the curve gives it the name' 'bath-
tub curve."
The first part of the curve shows a declining failure rate with time,
which could be described by a hyperexponential distribution. For this
form, the instantaneous failure rate is given by (Jardine, 1973)
r(t) = 2A{K2 + (l - K? exp[ - 2At(l- 2K)]
K + (l - K) exp[ - 2At(1- 2K)]
where A is the mean arrival rate of breakdowns with t ::::: 0, and K is a
parameter of the distribution (0 ::s K ::s 0.5).
The declining failure rate with time can be interpreted as some im-
provement in the equipment with use. It corresponds to the running-
80 6. Machinery Management

r(ll

Fig. 6.1. Bathtub curve.

in period for a new machine and would ideally occur within the ma-
chine's warranty period. The middle part of the bathtub curve shows
a constant failure rate with time. Failures in this region occur randomly,
and this period can be taken as representing the useful operating life
of the machine. The instantaneous failure rate ret) in this region can
be assumed to follow a negative exponential distribution
ret) = A
where A is the mean arrival rate (of breakdowns).
The latter part of the curve represents the period when the failure
rate is increasing. The machine or component can now be regarded as
being beyond its useful operating life. It would be replaced or given a
major overhaul. This aging or wearing out effect might follow a normal
distribution, or some other distribution indicating an increase in the
instantaneous failure rate.
The general definition of reliability is that a device will perform as
intended, without failure, for a specified period of time under given
operating conditions. In order to conduct an assessment of reliability,
the definition of failure, as well as the operating period and the op-
erating conditions must be given. For example, a failure need not be
just the condition that prevents operation of a machine-it could also
include those situations in which the machine's performance falls below
some acceptable level. The Weibull distribution is often used in failure
6.3 Reliability 81

rate studies because it is a three-parameter distribution and failure rate


patterns that are decreasing, constant, or increasing can be described
by varying the parameter values.
The Weibull cumulative density function is given by
F(t) = 1 - exp[ - (t - 'Y)I3/a]

where a is a scale parameter, 13 a shape parameter (i.e., it determines


whether the model has increasing, constant, or decreasing failure rate),
and 'Y a parameter representing the lower bound of the machine's life.
For agricultural machinery, the lower bound can be assumed to be
zero since the first failure can be expected as soon as the machine is
placed in service (Kumar et al., 1977). Thus 'Y = 0, and
F(t) = 1 - exp[ - (tJ3)/a]

This cumulative distribution function can be analyzed graphically by


the natural log twice to give a linear equation:
1
In In 1 _ F(t) = 13 In t - In a

lfthe data fit the Weibull distribution, then plotting In In {11[1 - F(t)]}
as the ordinate and In a as the abscissa will give a straight-line rela-
tionship. The parameters a and 13 can then be determined.
The data required are the times between failures. This can be plotted
as a histogram of number of observations and times between failures.
Kumar et al. (1977) used the Weibull distribution, developing equa-
tions for predicting the reliability of a combine harvester. They col-
lected data on six combine harvesters of the same make and model.
Calculated values of parameters a and 13 for the combine harvester
were
13 = 1.735, a = 961.609

Thus, the distribution predicting the time between failures for the com-
bine harvesters studied is given by
F(t) = 1 - exp[ - t1.735/961.609]

It should be noted that when 13 = 1, the failure rate is constant and


the Weibull distribution reduces to the exponential distribution. When
13 > 1, the failure rate is increasing with time.
82 6. Machinery Management

6.3.1 Series and Parallel Arrangements


The petformance of each component of a machine (or of each ma-
chine in a system) can be represented by a binary random variable Xi
such that (during some period 0 to t)
X =
I
{I for satisfactory petformance of component i
0 if breakdown of component i occurs

If <I>(X I , X 2, ... ,Xn) represents the petformance of a system, then


for a series arrangement of n components where the system will fail if
one component fails,
<I>(X I , X 2, ... , Xn) = X lX 2 ... Xn = min(X I , X 2, ... ,Xn)
Thus <1>(1, 1) = 1, whereas all other combinations of X I and X 2 give
zero, i.e.,
<1>(1, 0) = 0, <1>(0, 1) = 0, <1>(0, 0) = 0
For a parallel arrangement of n components, the system will only fail
if all the components fail. If one component functions, the system will
petform satisfactorily. The system petformance is given by
1 - (1 - Xd(1 - X 2) ... (1 - Xn)
max (Xl , X 2 , ... , Xn)
e.g., for two components Xl, X 2,

If Xl = 1, X 2 = 0,
<1>(1,0) = 1, <1>(0, 1) = 1, <1>(1, 1) = 1
The probability with which a series or parallel system of components
will operate can be expressed in terms of the above equations by con-
sidering the probability of the successful operation of an individual
component over the specified time period. If Pi is the probability of
successful operation of component i, then the reliability of the system
Rs is
n

Rs = PlP2P3 ... Pn IIpi


i~l

For a parallel system,


n
Rs = 1 - (1 - pd(1 - P2) ... (1 - Pn)
6.4 Selection of Machinery Capacity 83

It should be noted that 1 - Pi is the probability of failure of component


i.
If qi = 1 - Pi, then for a parallel system
n
Rs = 1 - II qi
i~1

If an operation is deemed to be critical relative to the overall perform-


ance of a system, then the operation can be duplicated to provide a
higher level of reliability. This constitutes having two series systems
in parallel, where Pi is the probability of the successful operation of
component i. In the event of breakdown of one operation, a second,
back-up operation is available.
The probability of failure of one operation is given by
n

The reliability of the two operations in parallel is the probability that


at least one will operate. This is given by (l - the probability offaiiure).
Thus, the system reliability Rs is given by

Rs = 1- (1- i~ Pir
The costs associated with reliability decisions can be incorporated into
the models by considering the economic penalty associated with a
breakdown, and the costs of providing back-up equipment.
Further consideration of reliability is beyond the scope of this text,
but the interested reader is referred to operations research texts such
as Hillier and Leiberman (1974), and also to those sections of such
texts that deal with queueing theory (waiting-line analysis). Many of
the problems associated with queueing theory, such as multiple-server
analysis of the repair facilities necessary to reduce machine downtime,
are applicable to machinery management decisions encountered in ag-
riculture.

6.4 SELECTION OF MACHINERY CAPACITY

The field capacity C (ha/hr) that is obtained from the operation of a


particular machine can be expressed as follows:
C = Swe/IO
84 6. Machinery Management

where S is the field speed (km/hr), w the implement width (m), and e
the field efficiency (decimal). The field capacity that is required can
be estimated by considering the time available for field work. This
capacity is given by
C = AlDhp
where A is the area to be worked (ha), D the number of days available,
h the expected number of hours available per day, and p the probability
of a working day.
This approach represents an initial starting point, but of more im-
portance is the machine capacity that represents the lowest net cost.

6.4.1 Hunt Model


Hunt (1964) developed an equation to select the optimum machine
width by expressing the purchase price of the machine as some unit
of its operating width. The optimum machine width w (m) was obtained
by setting the first differential of the annual cost equation to zero. The
equation obtained is as follows:
lOA ] 1/2
[
w = (FC%) pSe (L + T + KYvA)

where A is the annual area over which the machine is used (ha), FC%
the fixed-cost percentage (as a decimal) of the purchase price, p the
purchase price of the machine expressed per unit of effective operating
width ($/m) , S the forward operating speed (km/hr) , e the field effi-
ciency (decimal), L the labor cost ($/hr), T the tractor use cost ($/hr)
(T = 0 for self-propelled machines), K a linear timeliness coefficient,
and Yv is the value of the gross crop yield ($/ha).
The Hunt model provides a fairly simple approach to determining
the optimum width of machine. It assumes, however, that the purchase
price can be reasonably expressed on a per-un it-width basis, and uses
a straight-line depreciation method for determining the fixed costs.

6.4.2 Tullberg Model


Tullberg (1978) has suggested that the Hunt approach is unneces-
sarily complex and that a similar expression for optimum size can be
obtained by considering the capacity of the machine rather than its
6.4 Selection of Machinery Capacity 85

Annual
capital
cost

Fig. 6.2. Annual capital costs vs. machine


Capacity capacity [from Tullberg (1978)].

operating width. Tullberg breaks down the major costs of owning and
operating farm machinery into three categories: fixed, time, and task
costs.
1. Fixed costs are those related to ownership of machine capacity.
They are largely independent of utilization of the machine. Costs in
this cateogry include depreciation, interest, housing, and insurance.
They can be calculated by any of the accepted methods, and are ex-
pressed in cost ($) per annum.
2. Time costs are those related only to the time for which the ma-
chine is operating. Costs in this category including labor and penalty
costs associated with untimely operations ($/hr).
3. Task costs are those related to utilization or amount of work
carried out by the machine. Examples in this category include main-
tenance and repairs, and costs of consumables such as fuel ($/ha).
In this way, variable costs are divided between those related to the
time (time costs) and those related to the task (task costs). The annual
capital cost of machinery ownership might be linearly related to the
machine capacity, as shown in Fig. 6.2. Annual time costs will be
inversely related to machine capacity as shown in Fig. 6.3. Task costs

Annual
time
costs

Fig. 6.3. Annual time costs vs. machine ca-


Capacity pacity [from Tullberg (1978)].
86 6. Machinery Management

Total
annual
cost

Fig. 6.4. Total ownership cost VS. ma-


chine capacity [from Tullberg (1978)). Capacity

can often be assumed to be independent of machine capacity so that


total annual costs for the ownership and operation of a machine are
equal to the sum of task and time costs. This is illustrated in Fig. 6.4.
The least-cost (optimum) machine capacity required will be at the point
where annual time costs are equal to annual capital costs.
U sing an optimization procedure that sets the first differential of the
total annual cost to zero, the optimum capacity C (ha/hr) can be iden-
tified by the following equation:
C = YtUlf
where t is the time cost ($/ha), U the utilization (halannum), and f the
fixed cost per unit of machine capacity, ($ annum -\ ha - \ hr)
The time cost at the optimum capacity is given by
t = Cj/H
where H is the total operating time per annum (hr). Tullberg stated the
optimum capacity in marginal cost terms, as follows: "You buy larger
equipment until the fixed costs of additional capacity are equal to the
savings (in labor and crop losses) expected with the use of that addi-
tional capacity" (p. 135). The advantage ofthis approach is that it yields
a solution in terms of capacity, which can then be interpreted and
analyzed for different options, such as several smaller capacity ma-
chines. It also permits complex or relatively simple approaches to the
determination of the fixed and variable costs involved.

6.5 REPLACEMENT THEORY

In simple terms, a machine on the farm should be replaced when the


cost of retaining the existing machine exceeds the cost of acquiring a
6.5 Replacement Theory 87

new machine. In replacement theory terminology, the existing machine


is referred to as the "defender" and the new machine is referred to as
the "challenger." Although the concept of replacement theory is easy
to put into words, it is much more difficult to determine the associated
costs. This arises because it is necessary to compare the future ex-
pected costs of the existing machine with the projected costs of the
proposed replacement machine. In addition, the defender and chal-
lenger have unequal future lives; hence it is necessary to make com-
parisons based on annual costs to determine which option will be
cheaper on the average.

6.5.1 Holding-Cost Approach to Replacement


The use of the holding-cost concept in determining the optimum time
at which a machine should be replaced was proposed by Dunford and
Rickard (1961). This approach is based on a theory for secondhand
markets that had been proposed earlier (Fox, 1957).
The major costs associated with owning and operating a machine are
the initial outlay, the repair and maintenance costs, and the running
costs. The assumption is made that running costs can be disregarded
in the decision as to when to replace because the repairs and main-
tenance are reasoned sufficient to maintain the running costs at a con-
stant level. Thus the concern is with the capital and cumulative repair
costs. The sum of these two costs is termed the holding cost. Figure
6.5 shows holding cost (PH) for a machine whose initial and replace-
ment cost are assumed constant with time (OP). The repair costs are
assumed to increase with age of the machine, and thus the holding-
cost curve (PH) is of increasing slope. The holding cost associated with
keeping the machine to a given age can now be determined. For ex-
ample, at an age OS the holding cost is given by IS. Assuming no
secondhand or scrap value, the optimum time for replacement is given
by the tangent to PH drawn from the origin. In this example, the op-
timum time for replacement is given by OT and at this point the average
rate of holding cost per unit of life will be a minimum. That is RT/OT
will be the minimum ratio of holding cost to machine life.
An advantage of this approach is that it can be used to show the
minimum sum an owner must receive for an existing machine when
considering its replacement time (T). If replacement is considered at
time S, the owner must receive an amount sufficient to be restored to
the minumum operating line OG. For this case, this amount is given
88 6. Machinery Management

H G

Cost

o"'-____ ______'____
~

s T
Time

Fig. 6.5. Holding-cost curve [from Dunford and Rickard (1961))

by IJ. The same reasoning is applied to any replacement time that is


beyond the optimum point. Investment allowances could also be in-
cluded by including these on an annual basis in the capital costs of
owning the machine. The effect of such allowances is to induce the
farmer to replace machinery more frequently.
The simple model presented can also be used to take into account
technological innovations that may be present in the new machine. If
the same type of machine is being purchased, then the new machine
will depress the trade-in value of the obsolete older machine. Thus the
trade-in value received can be taken as encompassing the obsolesence
effect caused by the technologically superior new machine. In addition,
if a decision is made to replace the machine without receiving the trade-
in necessary to be restored to the minimum-cost curve, then the short-
fall represents the amount the owner is prepared to pay for the tech-
nological improvements offered by the new machine.

6.5.2 Group Replacement Theory


The replacement of a number of similar items at fixed intervals is
referred to as a group replacement policy. This approach often becomes
feasible if there are a large number of identical low-cost items that are
increasingly liable to failure with age, and when the value of the in-
6.5 Replacement Theory 89

dividual item is small (Ackoff and Sasieni, 1968; Riggs, 1978). Many
agricultural machines may be fitted with items that fall into this clas-
sification, for example, the nozzle tips on sprayer booms or the tines
on cultivators.
A group replacement model assumes that items that fail within a
given period are replaced individually, and that all items are replaced
at some interval, irrespective of when an individual item was previously
replaced. The system considered must be analyzed to determine (1) if
group replacement is worthwhile (Le., beneficial relative to individual
replacement), and (2) the appropriate group replacement interval that
minimizes the costs. The information required for such an analysis is
given as follows:
Let P k be the probability of a new item failing in the interval kt to
(k - 1)t, and let ik be the replacement at time kt. Initially, the system
is comprised of N new items (i.e., at t = 0), and at time nt all items
are replaced. The replacements made at time (k - 1)t will have a
fraction Po that fail by time kt; the replacements made at time (k -
2)t will have a fraction p\ that fail between (k - 1)t and kt; the re-
placements made at (k - 3)t will have a fraction P 2 that fail between
(k - 1)t and kt, etc. Thus

fk = f O P(k-1) + f\P(k-2) + ... + f(k-I)PO


The total number of replacements is given by summation as
n-\

L fi
k-\

The cost over n periods is


n-\

C[ L fi + foCo
k~\

where C[ is the cost of replacing an individual item and Co is the cost


of replacing all the items as a group. The cost per interval C is
1 n-\
Ct =- C[ L fi + foCo
n k~\

The period n must be chosen to minimize Ct. This is chosen as the


first minimum of the cost curve, provided this is less than the cost of
individual replacement. For this situation, the first minimum will be
90 6. Machinery Management

Cost

A ~---------------------------------
Time (n periods)

Cost

B ~-----------------------------------
Time (n periods)

Fig. 6.6. Group replacement analysis. A. Group replacement justified. B. Group


replacement not justified.

less than any other minimum. Figure 6.6A shows a cost curve that has
its first minimum below the individual replacement cost, showing that
periodic group replacement would be preferable to individual replace-
ment. Figure 6.6B shows a cost curve that does not justify group re-
placement.
The cost of individual replacement is determined from NC;/average
life. The average life is given by
Po + 2P 1 + 3P 2 + 4P 3 + ...
This cost must be compared with the first minimum cost resulting from
group replacement.
References 91

Group replacement theory provides for evaluation of replacement


cost so that the most beneficial decisions can be undertaken relative
to the replacement of low-cost similar items that are increasingly liable
to failure with age. The costs of individual and group replacement need
not be in monetary terms. They could also be expressed in terms of
the time taken for individual and group replacement.

REFERENCES
Ackoff, R. L., and Sasieni, M. W. (1968). "Fundamentals of Operations Research."
Wiley, New York.
ASAE (I 983a). "Agricultural Engineers Yearbook, 1982-83," D.230.3 Section 6.1, p.
218. Am. Soc. Agric. Engr. St. Joseph, MI.
ASAE (l983b). "Agricultural Engineers Yearbook, 1982-83," D.230.3 Section 3.3, p.
213. Am. Soc. Agric. Engr. St. Joseph, MI.
ASAE (l983c). "Agricultural Engineers Yearbook, 1982-83," D.230.3 Section 6.1, p.
219. Am. Soc. Agric. Engr. St. Joseph, MI.
Bartholomew, R. B. (1976). The economics of acquiring farm machinery. Tech. Bull. 9.
Department of Primary Industries, Queensland, Australia.
Dunford, W. J., and Rickard, R. C. (1961). The timing of farm machinery replacement.
J. Agric. Econ. 14, 348-358.
Fox, A. H. (1957). A theory of second-hand markets. Economics XXIV(94), May.
Hillier, F. S., and Lieberman, G. J. (1974). "Operations Research." Holden-Day, San
Francisco.
Hunt, D. R. (1964). "Farm Power and Machinery Management," 4th ed. Iowa State
Univ. Press, Ames, IA.
Hunt, D. R. (1973). "Farm Power and Machinery Management," 6th ed. Iowa State
Univ. Press, Ames, IA.
Jardine, A. K. S. (1973). "Maintenance, Replacement and Reliability." Pitman, Bath,
England.
Kumar, R., Goss, J. R., and Studer, H. E. (1977). A study of combine harvester reli-
ability. Trans. ASAE 20,30-34.
Ozkan, E., and Edwards, W. (1983). U sing microcomputers for machinery management.
ASAE Paper No. MCR 83-132. Am. Soc. Agric. Engr. St. Joseph, MI.
Pitt, R. E. (1980). Risk assessment as a decision making tool for agricultural engineers.
ASAE Paper No. NAR 80-102. Am. Soc. Agric. Engr., St. Joseph, MI.
Riggs. J. L. (1978). "Production Systems: Planning Analysis and Control." Wiley, New
York.
Tullberg, J. N. (1978). Farm machine capacity selection. Proc. 1978 Conf Agric. Engr.
pp. 134-137. Toowoomba, Queensland, Australia.
Von Bargen, K., and Cunney, M. B. (1974). Activity ratios for farm machinery operations
analysis. Trans. ASAE 17,225-227.
7
Vehicle Performance

The performance of road vehicles is primarily concerned with their


ability to negotiate grades, to accelerate, and to maintain a desired
road speed for a given load. The performance of these vehicles is ana-
lyzed by considering equations for motion and for tractive effort in
combination with the power plant and transmission characteristics.
The agricultural tractor, however, must be evaluated in terms of its
performance in generating drawbar power on differing soil surfaces and
with a range of implements. It must also be evaluated in terms of its
ability to control attached implements adequately. These are usually
attached to the rear of the tractor on a three-point linkage and are
hydraulically controlled. However, many tractors are also used with
hydraulically operated front-end loaders, and there is an increasing
tendency to offer a front-mounted three-point hitch as an option.

7.1 DRAWBAR PERFORMANCE

Drawbar power is given by


Po = FoV
where Po is the drawbar power (kW), Fo the drawbar pull (kN), and
V the actual speed (msec - I)
92
7.1 Drawbar Performance 93

The drawbar pull is given by


FD = F - ~R
where F is the tractive effort (kN) and ~R the sum of the resistance
forces (kN).
The drawbar power can also be expressed as
P D = (F - ~R)Vt(1 - i)
where V t is the theoretical forward speed (msec- I ) and i the slip (ex-
pressed as a decimal).
The resistance forces are due to
1. rolling resistance of the vehicle tires as they move along a given
surface,
2. flexure of the walls of the tire,
3. transmission resistance,
4. inertia resistance,
5. drawbar load,
6. air resistance,
7. grade resistance.
The determination of rolling resistance is discussed in Chapters 4
and 5. Becker (1969) has provided a method of incorporating the re-
sistance associated with flexure of the walls of the tire into equations
predicting the rolling resistance-the pressure due to the carcase stiff-
ness is added to the inflation pressure. This is compared with the nor-
mal pressure at the lowest point of contact between the tire and the
ground. If the sum obtained is greater than this normal pressure, the
tire is assumed to act as a rigid wheel, and the Becker equations for
motion resistance of a rigid wheel apply. If the sum is less than the
normal pressure, the tire sinkage is recalculated and a second equation
used to evaluate the rolling resistance due to ground compaction. The
resistance due to flexure of the tire walls is usually small and can be
neglected. The reader interested in its consideration should consult
Becker (1969) or Wong (1978).
The transmission resistance represents the energy consumed in de-
livering power through the transmission system components. It is not
usually expressed as an equivalent linear force, but is incorporated into
the value for the overall transmission efficiency. It is important to
recognize, however, that if the rolling resistance of a tractor is mea-
sured by the common method of towing it in neutral, then there will
94 7. Vehicle Pecformance

be an overestimate of the rolling resistance because of the transmission


resistance. This error can be minimized by towing the tractor at a low
speed (McCarthy and Kolozsi, 1974).
The inertia resistance is the force that opposes the vehicle's change
of speed. For a vehicle of mass rn and a known rate of change of speed,
the inertia resistance RI is given by
RJ = rn dv/dt = rna
Change in forward speed of the vehicle is accomplished by a change
in rotational speed of the wheels and transmission system. The torque
M required for this change in rotational speed is given by
M = f dw/dt = fa

where f is the moment of inertia about the axis of rotation.


A vehicle has several groups of parts rotating at different speeds;
however, the torques necessary to achieve a change of speed can be
determined by considering the transmission ratio at some common
point:

where Mi is the torque required to produce acceleration a of all the


rotating components and ~ the reduction ratio between the common
point and the rotating component.
For example, if the common point is the drive axle, then
and
where ~ is the reduction ratio between the drive axle and the particular
component. Substituting for a gives
Mi = ad Lf~a
where ad is the acceleration at the drive axle.
The total moment of inertia is the sum of that due to translatory
motion of the vehicle and that due to rotation of the components. The
equivalent moment of inertia of the individual components at the fly-
wheel provides a convenient method of representing the inertia effects.
This is given by

where ~o is the gear ratio between the flywheel and the component i.
7.1 Drawbar Performance 95

The kinetic energy of a vehicle is given by


1 1
E = 2mv2 + 2"2:)w 2

where w is the angular velocity of the rotating components. The inertia


resistance RJ is given by
RJ = P - LR
where P is the tractive force and 'i.R the sum of all other resistance
forces.
Differentiating the equation for kinetic energy gives
dE = mv dv + LIw dw
The inertia resistance is given by

RJ = dV(m + LIe)
dt r2
where r is the rolling radius of the tire.
The term 'i.Ie!r represents the equivalent mass of the rotating com-
ponents assumed to be located at the rolling radius r of the vehicle
drive tires. This equivalent mass can be added to the vehicle mass to
give the total equivalent inertia mass:
m l = m + me
where m is the effective inertia mass and me the equivalent mass. The
l

equivalent mass me is obtained from


mi = mer ad = ad LIe
and so
me = LIe!r
For convenience, a term 'Y relating the relationship between m + me
and m is introduced, such that
'Y = (m + me)!m
and so
m l = m'Y
Thus

and therefore

'Y
96 7. Vehicle Performance

where Iw is the moment of inertia of the wheels and Ii the moment of


inertia of any part i rotating at engine speed with speed ratio S with
respect to the driving axle (Taborek, 1957a). The value of'Y can also
be obtained from the following empirical relationship, which is a modi-
fication of that provided by Wong (1978):

'Y = 1 + (~~~ + 0.0025e)

The total inertia resistance of the vehicle RIT is then given by


RIT = 'Yma

where m is the vehicle mass and a the vehicle acceleration.


The aerodynamic resistance of an agricultural tractor is of little in-
terest, particularly when the vehicle is operating in the field at low
forward speeds. However, the aerodynamic resistance is proportional
to the square of the velocity of the vehicle relative to the wind. Thus
in transport mode, when the tractor is likely to be traveling at speeds
up to 30 km hr- I , it becomes of some significance.
The aerodynamic resistance force Ra (N) can be expressed as

Ra = !CDPV/ A
where CD is a dimensionless drag coefficient, p the air density (kg m - 3)
Vr the velocity of the body relative to the air (m sec - I ), and A the
projected surface area of the body in the direction of travel (m2 ).
The drag coefficient can be determined by using scale models in wind
tunnels. The drag force for a particular vehicle shape and air velocity
can then be determined and thus a value for the drag coefficient ob-
tained.
A general prediction equation for the drag force is as follows (Mur-
phy, 1950):
Ra/p V2(l = f(xll, R)
where I is a length dimension, x other pertinent lengths, and R the
Reynolds number.
However, in order to maintain scaling of the Reynolds number, the
required model air velocity is usually extremely high. The Reynolds
number remains fairly constant at the low velocities encountered for
vehicle applications, and its effects can be ignored.
7.3 Vehicle Speed 97

The prediction equation for aerodynamic resistance force can then


be stated as
V2 n2
Ra = V~ Ram
where V m is the velocity of the air for the model, n the length scale
factor, and Ram the aerodynamic resistance of the model.
Wong (1978) gave values for Co for several types of vehicles. Pas-
senger cars have values in the range 0.3-0.6, a bus has values 0.6-0.7,
and a truck has values 0.8-1.0.
The resistance due to the effects of climbing a slope is determined
by considering the component of the vehicle's weight acting down the
slope. This is given by
RG = W sin e
where RG is the grade resistance, W the weight of the vehicle, and e
the angle of the slope in degrees.
The total forces resisting vehicle motion are the sum of those due
to drawbar load, rolling resistance, air resistance, grade resistance,
etc.

7.2 ENGINE TORQUE REQUIREMENT

Thrust-slip relationships for the traction of a vehicle tire were pro-


vided in Chapter 4. The engine output torque Me (Nm) that is required
to meet a thrust requirement depends on the value of the thrust, the
rolling radius of the drive tires, and the overall gear ratio:
M _ Fr
e - £01")1

where F is the thrust (N), £0 the overall gear ratio, 1")1 the overall trans-
mission efficiency, and r the rolling radius of the drive tires (m).

7.3 VEHICLE SPEED

The vehicle speed can be determined from the engine speed, the
overall gear ratio, the rolling radius of the drive tires, and the wheel
slip:

V = Nr(1 - i)
£0
98 7. Vehicle Performance

where V is the actual speed of the vehicle (m sec - I) and N the engine
speed (rads sec - I)

7.4 VEHICLE ACCELERATION

The torque required for the acceleration of a rotating part is given


by

where Mp is the torque acting on the rotating part, Ip the moment of


inertia of the part about its rotational axis, and Up the angular accel-
eration of the part.
For the vehicle as a whole, the acceleration a can be obtained from

where MA is the torque available for acceleration and I tot the total
moment of inertia of the vehicle.
If the moment of inertia is referred to the engine flywheel and the
torque available for acceleration is that provided by the engine, then
ae = dNldt = MAlltot E

where a e is the acceleration of the engine output shaft, N the engine


speed, and I tot E the total inertia of the vehicle referred to the engine
output shaft.
The load inertia Ie (kg m2 ) reflected to the engine output shaft can
be predicted from

NL)2
Ie = h ( Ne Ie = h
(1 )2
~o

where h is the inertia of the load (kg m2 ), NL the load speed, Ne the
engine speed, and ~o the overall gear ratio between the engine and the
load.
The total inertia (reflected to the engine output shaft) is given by
I tot = translatory inertia of the vehicle + inertia of the wheels
+ inertia of the rotating components + inertia of the engine.
In a computer simulation model, Dwyer et al. (1974) ignored the
7.5 Longitudinal Stability 99

inertia loads imposed by the rotating transmission components and


used the following equation:
W Ll V 2 2IRw 2h w Vt
I tot = gN2 + To
+ N 2?- + IE
where W is the vehicle weight (N), Ll V the change in theoretical forward
speed (m sec - I), g the acceleration due to gravity, N the engine speed
(rad sec-I), I Rw the inertia of the rear wheel (kg m2 ), I Fw the inertia
of the front wheel (kg m2 ), r the rolling radius of the undriven (front)
wheels (m), and IE the inertia of the engine (kg m2 ).

7.5 LONGITUDINAL STABILITY

The longitudinal stability of tractors becomes of particular impor-


tance when loads are carried outside the wheel base. Tractors are often
used with front- (and rear-) mounted loaders, and we now consider the
dynamic forces that result from such operations.
The static moments to be considered are those due to the load carried
in the vehicle and those of the vehicle alone. The other moments result

Q sin (j + Q a....-------
9

Fig. 7.1. Tractor and loader stability.


100 7. Vehicle Performance

from the acceleration of the vehicle in the drive direction and the ac-
celeration of the loader bucket in lifting or lowering the load.
It is useful to consider the worst possible case. This results from the
situation where the vehicle is being decelerated on a downward slope
and the load is being decelerated as it is lowered from an elevated
position. This is illustrated in Fig. 7.1. The acceleration of the vehicle
is a and the acceleration of the load is al. The height of the center of
gravity of the vehicle above the surface is h, and the height of the
center of gravity of the load carried by the bucket is hi. The other
pertinent dimensions are shown in Fig. 7.1.
For the balance condition, taking moments about point F gives

LfW cos e= LqQ cos e+ hi Q sin e + fJg ahl


·
+ h W sm W h +
e + -a Q
-alLq
g g
A stabilizing factor IT was introduced by Taborek (1 957b) in analyzing
the longitudinal stability of forklift trucks. This was defined as
stabilizing moment
IT = overturmng
. moment
If the previous balance equation is rearranged into a form such that:
1. The static level moments are given by

2. The moments due to slope are

B = tan e(Wh + Qhl)


WL f

3. The moments due to vehicle acceleration are

c = a(Wh + Qhl)
gLfW cos e
4. The moments due to load acceleration are
7.6 Tractor Dynamics 101

Then the stabilizing factor (J can be expressed as

1
(J = -------------
A+B+C+D
Note that on level ground, the stability factor is given by
(J = LfWILqQ
This analysis allows the longitudinal stability to be analyzed in terms
of the static and dynamic forces acting. An acceptable value for the
load lifted on an arm extending beyond the wheelbase can be deter-
mined from the stability factor.

7.6 TRACTOR DYNAMICS

Figure 7.2 shows a two-wheel drive tractor operating on a slope. For


the purposes of these analysis, it will be assumed that the front wheels
are relatively light, and that their inertia effects are negligible. The rear
wheels, however, have considerable inertia effects, and the analysis
of the dynamic forces acting is best conducted by considering the two-
wheel drive tractor as comprising one system involving the rear wheels
and axle, and a second system consisting of the body of the tractor,

Fig. 7.2. Tractor on a slope.


102 7. Vehicle Performance

Fig. 7.3. Rear axle subsystem.

the engine and the front axle. Figure 7.3 is a free-body diagram of the
rear axle subsystem.
The inertia coordinates are parallel (x) and normal to the plane (Y),
as shown in Fig. 7.3. For the subsystem consisting of the rear wheels
and axle, the equations of motion are:
W
~
g
x= F - TF - W w sin e- H (7-1)

W
~
g
y = N + W w cos e- R (7-2)

Iw ~ = M - (F - TF)r - Re (7-3)
From Fig. 7.3,
H
Hr = Re or e = -r
R
However, HIR = HIW = fL (coefficient of traction), and so
Iw~ = M - (F - TF)r - H
Since F - TF = H, then
Iw~ = M - H(r - 1) (7-4)
7.7 Tractor-Implement Forces 103

For the body of the tractor, the equations of motion are as follows:
W
H = Wb sin e- (TF)f - P cos ex = ~i
g
(7-5)

W B ..
W B cos e+ P sin ex - N - Rf = - y
g
(7-6)

where P is the drawbar pull acting from the implement onto the chassis
of the tractor. Taking moments about the center of gravity of the body
of the tractor,
RAL - Lr) + M - NLf + H(h - r)
+ P (sin ex)C - P(cos ex)h/ = 0 (7-7)

A more detailed approach, which considers, in addition, the effects of


the moment of inertia of the tractor body about its center of gravity
owing to a given pitch angle, is described by Liljedahl et al. (1979).
Although six degrees of freedom can be identified by considering the
two subsystems, the rear wheels and axle, and the tractor body, the
tractor as a whole has four degrees of freedom, and thus the tractor
dynamics can be described by four independent differential equations.

7.7 TRACTOR-IMPLEMENT FORCES

In the first case, we shall consider the performance of a three-point


linkage mounted implement, with no hydraulic system for implement
control. With this arrangement, when the implement is in work, it will
rise or fall relative to the tractor until the resultant of all soil and grav-

vhp

Fig. 7.4. Unrestrained three-point linkage controls.


104 7. Vehicle Performance

itational forces acting passes through the virtual hitch point (vhp). This
resultant is denoted by R in Fig. 7.4. In this situation, all external forces
on the implement will be in equilibrium. If, at any time, the resultant
R passes above the vhp, the implement will penetrate into the ground.
Adding weight to the implement will cause R to be steeper, thereby
increasing the distance from the line of action of R to the vhp. This
will have the effect of increasing the rate of penetration. Equilibrium
conditions are restored when the implement penetrates to such a depth
that R passes through the vhp.
With hydraulic pressure, the linkage no longer acts as a free-link
system, and the forces in the lower links cannot be assumed to pass
through the vhp. However, when the implement is being lowered, the
hydraulic system pressure is usually negligible, and the forces in the
links intersect at the vhp (Cowell and Len, 1967).

7.8 TRACTOR-IMPLEMENT DYNAMICS

The dynamic performance of a tractor-implement combination is


affected by the operational characteristics of (Dwyer, 1969) (1) the
engine, (2) the tire-ground interface, and (3) the implement depth or
draft control system.
1. The dynamic performance of the engine and transmission can
be described by the equation

I = dN + Hr = fl(N) (7-8)
dt Tj~

where I is the effective polar moment of inertia of the rotating parts


(kg m2 ), N the engine speed (rad sec-I), H the tractive force, r the
rolling radius of the drive wheels (m), Tj the transmission efficiency, ~
the overall transmission ratio, and fl a function obtained from the
steady-state torque-speed relationship.
2. The performance at the tire-ground interface can be described
by the following equations:
(7-9)
where Wa is the weight on the drive tire axle, i the slip of the drive
tires, and f2 a function describing the relationship between coefficient
of traction and slip; and P = H - p(WT + WI + Sy) where P is the
7.8 Tractor-Implement Dynamics 105

drawbar pull (N), p the coefficient of rolling resistance, WT the weight


of the tractor (N), WI the weight of the implement (N), and Sy the
vertical soil forces acting on the implement (N).
The forward motion of the tractor is described by

WT + WI dV =P _ D
g dt
where V is the actual forward speed of the tractor (m sec-I) and D
the implement draft force (N)
3. The dynamic performance of the tractor-implement depth or
draft control system was described by Dwyer (1969). The draft force
is given by
D = Qd
where Q is the draft force per unit of implement depth (Nm - I) and d
the depth (m). Thus, in this analysis, the draft force is assumed to be
proportional only to the depth of the implement.
The soil vertical forces acting (Sy) are the sum of the steady-state
and dynamic soil forces. The steady-state soil force can be defined as
the mean vertical force at a certain depth. The dynamic vertical soil
force component is only present when the implement is moved verti-
cally relative to the soil surface. Thus
dy
Sy = Qd dx + Rd
where R is the increase in steady-state vertical soil force with depth
(Nm- I), y the vertical movement relative to the ground, and x the
forward movement.
Figure 7.5 shows a three-point mounted implement. For top-link sen-
sing, the sensed force T is obtained by taking moments about point X:
Db SyC Wle
T=-----
a a a
where T is the horizontal component of the top-link force and D the
draft force. Thus

T = Pdb _ (PddY + Qd)£ _ WIe


a dx a a
106 7. Vehicle Performance

W,

Fig. 7.5. Forces acting on a three-point hitch for implement control.

where d is the depth. The depth can be represented in terms of the


mean depth d m and the depth error de, such that d = d m + de. Thus

T = P(dm + de)b - [P(d m + de/dY + Q(d m + d e )]£ _ W1e


a x a a
The set top-link force To is given by
Pdmb QdmC W1e
T=-------
a a a
The sensed-error signal is then given by I(T - To)1 and the draft control
will produce a rate of vertical movement of the implement that is pro-
portional to this sensed error.
The analysis of the lower-link sensing system is provided in a similar
manner (Dwyer, 1969). Taking moments about the top-link implement
pin,
D(a + b)
L
a a a

where L is the sum of the horizontal components of the forces in the


lower links.
The sensed force is X (Fig. 7.6), which is the horizontal component
7.8 Tractor-Implement Dynamics 107

Fig. 7.6 Lower-link sensing.

of the forces acting at the front fixing pins of the lower links. Resolving
forces in the horizontal direction gives
x + R cos e = L
where R is the force acting in the lift rods.
Taking moments about the front fixing pins,
R(sin e)j cos <I> - R(cos e)j sin <I> = (WI + Sv) h cos <I> - Lh sin <I>

and so

R = ..:....(W~I_+_S..:....v)_h_c_o_s_<I>'-..--_L_h_s_in---'..<I>
j sin(e - <1»

The sensed force X is given by

X = L (WI + Sv)h cos <I> - Lh sin <I> '"


cos~
j sin(e - <1»

Dwyer (1969) simplified this equation by assuming that angle <I> would
be relatively small and could be assumed to be zero. Also, angle e was
assumed to be a constant.
108 7. Vehicle Performance

The sensed force X is then given by


X = L _ (WI + Sv)h
j tan e
or
D(a + b)
X
a

REFERENCES
Becker, M. J. (1969). "Introduction to Terrain Vehicle Systems." Univ. of Michigan
Press, Ann Arbor, MI.
Cowell, P. A., and Len, S. C. (1967). Field performance of tractor draught control sys-
tems. 1. Agric. Engr. Res. 12(3), 305-221.
Dwyer, M. J. (1969). The dynamic performance of tractor-implement combinations.
Proc. Inst. Mech. Engr. 184(3Q), 68-82.
Dwyer, M. J., Crolla, D. A., and Pearson, G. (1974). An investigation of the potential
for improvement of tractor draught controls. 1. Agric. Engr. Res. 19, 147-165.
Liljedahl, J. B., Carleton, W. M., Turnquist, P. K., and Smith, D. W. (1979). "Tractors
and Their Power Units," 3rd ed. Wiley & Sons, New York.
McCarthy, T. T., and Kolozski, Z. (1974). The measurement of tractor transmission
losses. 1. Agric. Engr. Res. 19,71-75.
Murphy, G. (1950). "Similitude in Engineering." Ronald Press, New York.
Taborek, J. J. (l957a). Vehicle resistance forces. Machine Design, August 8, 101-107.
Taborek, J. J. (1957b)) Longitudinal stability. Machine Design, September 5,109-114.
Wong, J. Y. (1978). "Theory of Ground Vehicles." Wiley & Sons, New York.
8
Connecting Tractor and
Implement*

The working unit for both field work and transportation consists of a
tractor and the implement, which are connected together by compo-
nents which are called the hitches. The tractor and the implement in-
fluence each other with regard to function, and consequently the
method of connecting them has important influences on the effective-
ness of the entire working unit.
Basically the function of the hitching elements is to transfer forces
between implement and tractor and to govern the movement and po-
sition of the implement relative to the tractor. A third function of the
hitch components is to permit replacing one implement by another.
When selecting or designing a hitching system the following factors
should be considered:
1. Maximum efficiency in power utilization should be achieved for
the entire working unit.
2. Switching from one implement to another should be performed
as easily as possible.

* This chapter was written by Sverker Persson, Agricultural Engineering Department,


Pennsylvania State University, University Park, PA.

109
110 8. Connecting Tractor and Implement

3. The hitching system should ':Je effective both during work and
during transport.
4. The hitching system should be capable of accommodating a wide
range of implements on the same tractor.

8.1 PULL-TYPE, SEMIMOUNTED, OR MOUNTED


IMPLEMENT

With regard to hitching, the implements may be of the pull type (Fig.
8.1), fully mounted (Fig. 8.2), or semimounted (Fig. 8.3). These are
the most common methods of hitching the implements but for special
implements other forms of mounting are also used. The difference be-
tween the three methods is basically a difference in the vertical forces
that are transferred between the implement and the tractor and in the
amount of control that the tractor operator has on the implement. Pull-
type implements always have means for supporting their weight and
other vertical forces. A fully mounted implement does not have any
support devices of its own and depends entirely on the tractor for
balancing vertical forces. The semimounted implement is partially sup-
ported by its own wheels and partially by the tractor.
The selection of the type of hitching is influenced by the type of
implement and the size of implement and tractor. Normally, fully
mounted implements are used on small tractors, where the implement
is close to the tractor and therefore can be effectively supported by
the tractor. Large implements have the center of gravity so far behind
the tractor that they cannot practically be carried by the tractor.
Because relatively smaller vertical forces can be transmitted between

Fig. 8.1. Pull-type implement.


8.2 Hitch Point 111

Fig. 8.2. Fully mounted implement.

Fig. 8.3. Semimounted implement.

a pull-type implement and the tractor than between a mounted imple-


ment and the tractor, a tractor built for pull-type implements usually
must be heavier than a tractor built for mounted implements. The spe-
cific calculations related to this transfer of vertical load wiII be dis-
cussed in more detail later.

8.2 HITCH POINT

When the implement and the tractor are connected in only one point
and this point is flexible in all directors, the resultant of all forces
transferred between the implement and the tractor has to pass through
this point at all times. This defines the hitch point for the system (Fig.
8.4).
The hitch point also determines how the implement wiII turn relative
to the tractor during turns or when passing over uneven ground. The
112 8. Connecting Tractor and Implement

Relative motion

Fig. 8.4. Single-point hitch.

hitch point consequently is of significance both for the force transfer


and for the relative motion between implement and tractor.
When the implement and tractor are connected at three or less points
that are not flexible, these actual hitch points transmit several force
components and also moments between implement and tractor. In
some cases, the design of the attaching members permits determination
of one point in the system through which the resultant forces between
implement and tractor must pass. This is called the virtual hitch point
(vhp). Sometimes the vhp also determines the relative motion between
implement and tractor. In the analysis of the system it consequently
becomes a question of determining the position of the vhp and thence
determining the force transfer between the two parts (Fig. 8.5).

vhp

Fig. 8.5. Three-point hitch.


8.3 Load Transfer Due to Pull 113

8.3 LOAD TRANSFER DUE TO PULL

The supporting forces balancing weight and other vertical forces in


the tractor-implement system are located at three points: the front
axle, the rear axle, and the implement.
The magnitude of these supporting forces can be determined for a
stationary tractor by putting it on a scale. The values determined in
this way are called the static loads on front axle, rear axle, and im-
plement.
As soon as the tractor moves and pulls an implement with a resis-
tance, a change in these supporting forces will occur, which is called
the load transfer. The amount of load transfer is very much influenced
by the method of hitching the implement to the tractor. The loads when
the tractor is moving and pulling are called dynamic loads. Load trans-
fer is the difference between the dynamic load and the static load at
each respective support location.
Under normal working conditions, with the implement behind the
tractor, load transfer occurs from the front axle to the rear axle of the
tractor and from the implement to the rear axle of the tractor. In other
cases, as for instance with a front-end loader on the tractor or front-
mounted implements, load transfer could occur in the opposite direc-
tion.
The magnitude of the load transfer can be calculated using moment
equations (see Fig. 8.6). Figure 8.6A shows the static situation with the
tractor standing still and not pulling. The supporting forces from the
ground are located directly below the centers of the front and rear
wheels and at some point in the implement. As soon as the tractor
moves forward, two things happen: (1) the supporting forces from the
ground on the tractor wheels move ahead of the center of the wheel
as shown in Fig. 8.6B, and (2) a force between the tractor and the
implement is created through the hitch point due to the draft from the
implement.
The static support forces can be calculated using the notation shown
in Fig. 8.7 by writing one force equilibrium equation and one moment
equation:
Force equilibrium: N 2 (s) + N,(s) - W = 0 (8-1)
Moment equilibrium: Nl(si - Wx, =0 (8-2)
For the dynamic case (Fig. 8.8), the effects of motion resistance,
114 8. Connecting Tractor and Implement

Fig. 8.6. A, Static; B, dynamic reaction forces.

pull (Fo ), slope (8), and forward acceleration (i) are added. The force
and moment equilibrium equations become
Wi
- Wx - - + H - R - Fo(x) = 0 (8-3)
g
- Wy + N2 + N\ - Fo(y) = 0 (8-4)

- N 2(x\ - X2) - (Wx + W') gX y\ - N\(X3 - X2) (8-5)

- F O(x)Y2 - F o (Ylx4 + X2) + Wy(x\) = 0


where W is the weight of the tractor acting from the center of gravity,
Wx the component of the tractor weight acting in the x plane, Wy the
8.3 Load Transfer Due to Pull 115

Center of gravity

Fig. 8.7. Static support forces.

component of the tractor weight acting in the y plane, i the vehicle


acceleration in the x plane, H the thrust developed, R the rolling re-
sistance, FD the drawbar pull (with FD(x) and FD(y) representing the
components of the drawbar pull acting in the x and y planes, respec-
tively), NJ the reaction at the front wheels, and N z the reaction at the
rear wheels.

Fig. 8.8. Dynamic forces acting.


116 8. Connecting Tractor and Implement

The linear dimensions XI, X2, X3, X4 and YI, Y2 are as defined in Fig.
8.8.
The distances X2 and X3 can be calculated from the laws for motion
resistance (rolling resistance). The difference X3 - X2 is often very
nearly equal to the wheelbase, t.
The selection of positive sign for the variables in Eqs. (8.3)-(8.5)
corresponds to the most common directions of these vectors for a rear-
wheel driven tractor. Attention must be given to the signs when using
Eqs. (8.3)-(8.5) for cases other than those shown.
The load transfer is, by definition,
LT2 = N2 - N 2(s) (8-6)
LTI = Nl(s) - NI (8-7)
where LT2 is the load transfer to rear axle, LTI the load transfer from
front axle, N 2(s) the static rear wheel reaction, NI(s) the static front
wheel reaction. The load transfer may be calculated combining Eqs.
(8.1)-(8.7).

8.4 EFFECT OF LOAD TRANSFER

For a rear-wheel driven tractor, the load transfer as calculated above


benefits the performance of the tractor in the following ways:
1. The vertical load on the powered rear wheels is increased, per-
mitting an increase in the net traction without increase in wheel slip.
2. The motion resistance on the front axle is reduced because the
load on the front wheels is reduced.
3. As a consequence, the available drawbar pull is increased.
If a possible increase in drawbar pull is not required, the effect will
instead be a lower required coefficient of net traction and consequently
a lower travel reduction for the drive wheels.

8.4.1 Limits for Load Transfer


The amount ofload transfer is limited by two factors: tractor stability
and implement function.
If the load transfer from the front axle is of such magnitude that only
a small portion of the tractor weight is left on the front axle, the tractor
8.4 Effect of Load Transfer 117

Fo
~------~~~----------~~T-----------

Fig. 8.9. Load transfer.

becomes difficult to steer. If the load transfer from the front end is
equal to the static load on the front axle, the tractor becomes unstable
and may tip over backward or at least the front wheels may rise up in
the air. Both conditions are undesirable, and a limit should therefore
be set on the load transfer from the front wheels such that 10-20% of
the tractor weight remains on the front wheels at all times.
If the load transfer from the implement is very high, it may be difficult
for the implement to maintain a constant working depth or perform its
desired function.

8.4.2 Factors Influencing the Load Transfer


As evident from the previous calculations, the load transfer is influ-
enced by the following factors:
1. The magnitude of the draft force F D.
2. The distance from the draft force reaction line to the rear-wheel
supporting point relative to the wheel base, i.e., as the ratio hll (Fig.
8.9). This distance will increase if the hitch point is high above the
ground, if the hitch point is far behind the rear axle with the force
pointing downwards, and if the draft force has a considerable vertical
component.

8.4.3 Static Load Distribution for Two- and Four-


Wheel Drive Tractors
Theoretically the most desirable condition when maximum traction
is required would be that all available vertical forces be supported by
118 8. Connecting Tractor and Implement

70%
A

B
Fig. 8.10. A, Static-load; B, dynamic-load distribution, two-wheel drive.

powered wheels. For a rear-wheel driven tractor this means that the
load should be concentrated on the rear wheels and only the necessary
minimum left on the front wheels for stability and steering. Static load
distribution between front and rear axle, as shown in Fig. 8.10A, is then
desirable. The dynamic load distribution is also shown in Fig. 8.lOB.
A consequence of this arrangement is that only a little additional pull
can be produced by powering the front wheels on a tractor built with
this type of load distribution, since little vertical load remains on the
8.4 Effect of Load Transfer 119

40% 60%
B
Fig. 8.11. A, Static; B, dynamic load distribution, four-wheel drive.

front wheels when the tractor is pulling under normal conditions. This
fact has been established through practical tests.
In order for this amount of load transfer to occur, the draft force
must be located as shown (Fig. 8.9) and be of such a magnitude that
it gives the required load transfer. Under some circumstances, this does
not happen. Consequently it may be justified to provide power also to
the front wheels for normally rear-wheel driven tractors. Examples of
such cases are as follows:
1. When the load from the implement comes in front of the tractor,
as for instance when using a front-end loader: In this case, considerable
load transfer occurs from the rear to the front axle.
120 8. Connecting Tractor and Implement

2. When driving conditions are difficult with high coefficient of


motion resistance and low coefficient of net traction: In this case, the
magnitude of the draft force is low and the magnitude of the motion
resistance on the front wheels is high. The difference between the two
will be considerably improved if the motion resistance of the front
wheels can be changed to provide some tractive force.
3. When working on steep, sloping terrain, where a considerable
load transfer between front and rear axle occurs due to the slope. Under
such circumstances safety considerations require that the static front-
end load should be fairly large, making powering of the front wheel
more justified.
4. When the required load on the rear wheels for full utilization of
the engine power would be larger than desired with regard to suitable
tires or soil compaction: In this case the weight of the tractor would
have to be distributed over more tires including the front wheels. This
is the case for the regular four-wheel drive tractors with equal-sized
tires front and rear. Such a tractor would have a static and dynamic
load distribution as shown in Fig. 8.11.

8.5 FREE-LINK THREE-POINT HITCHES

Originally, three-point hitches were made such that the hydraulic


power was used only when the implement was raised out of the working
position for transport. During working no hydraulic power was applied.
The forces in the three links would then have to act along the links
because the links had more or less frictionless joints at both ends. In
this way a vhp could be established as shown in Fig. 8.12. As can be
seen, this vhp is far ahead of the actual hitch point (ahp) used for pull-
type implements. A fairly small lo~d transfer from the front wheels
would occur even if the draft force and the load transfer from the
implement were considerable.
If the links of this hitch were adjusted such that they were almost
parallel, the vhp would be far ahead of the tractor. The motion of an
implement relative to a tractor will be almost parallel up and down,
which is desirable, for instance, for a cultivator.
For a plow, the vhp should be closer to the rear of the tractor in
order to make the plow self-adjusting in depth.
8.6 Powered Three-Point Hitch 121

Fig. 8.12. Free-link vhp.

8.6 POWERED THREE-POINT HITCH

If the operator wants to increase the amount of load transfer or to


be able to control the implement according to working conditions, hy-
draulic power could be used to apply a lift force in the lower links of
the three-point hitch. The force diagrams are shown in Fig. 8.13.
The maximum load transfer that could take place from the implement

~------ X6 --------~

Fig. 8.13. Constrained three-point linkage.


122 8. Connecting Tractor and Implement

Fig. 8.14. Vertical movement of a mounted implement.

to the tractor in such a system as far as it is limited by the hitch is


calculated from the following equation (Fig. 8.13):
WI = LXS/X6

where WI is the weight of the implement, L the lift force at the end of
the link, X6 the distance of the load behind the vhp, and Xs the distance
of the end of the link behind the vhp.
It would be noticed that if the vhp is located far ahead of the tractor,
the implement load may be located far behind the tractor and the hy-
draulic system will still be able to lift it. This will make the tractor
unstable and consequently require an improvement of the tractor sta-
bility by added front-end ballast.

8.7 MOTIONS OF FULLY MOUNTED IMPLEMENT


IN THREE-POINT HITCH

An implement mounted in a three-point hitch as shown in Fig. 8.14


will move relative to the tractor as if it were turning around the vhp.
That means that if the implement moves up relative to the tractor it
will, at the same time, be tilted more forward, and if it moves down
relative to the tractor it will be tilted more backward. This fact is uti-
lized to give an implement, for instance, a moldboard plow, a self-
adjusting motion. The plow will tend to return automatically to a po-
8.9 Draft-Sensing Hydraulic Systems 123

- -:: .....

Fig. 8.15. Lateral movement of three-point hitch.

sition where its bottom is parallel to the forward motion and the land
surface.
In addition, the three-point hitch may give a self-adjusting motion
to sideways movement, as shown in Fig. 8.15.
This self-adjusting effect naturally becomes more pronounced if the
vhp is closer to the implement. If it is far out ahead of the tractor, there
is no self-adjusting tendency and the implement gets a parallel motion
instead.

8.8 LIFTING CAPACITY OF THE HYDRAULIC


SYSTEM

As mentioned above, the ability of a hydraulic system to lift an im-


plement or a load is influenced by the distance of that implement or
load both to the ends of the lift arms and to the vhp. In order to stand-
ardize the specifications for the lifting capacity L of a hydraulic sys-
tems, ASAE (1983) has established relevant standards.
Due to the movement of the vhp and of the linkage in the hydraulic
system, the lifting capacity of the hydraulic system normally varies
with lift height.

8.9 DRAFT-SENSING HYDRAULIC SYSTEMS

In order fully to utilize the load transfer capabilities of the three-


point hitch, it is necessary to adjust the load transfer such that at each
moment the maximum load transfer takes place. The working maxi-
124 8. Connecting Tractor and Implement

mum is exceeded when the implement is lifted and does not maintain
proper working depth. Variations in the ability of the implement to
permit load transfer happen very fast in field operations due to the
uneveness of the soil; consequently it is not possible to adjust the load
transfer manually, but an automatically controlled load transfer system
has to be incorporated.
One of the earliest and still a fairly common system for automatically
controlling the working depth is the draft-sensing system. Instead of
measuring or sensing the working depth of the implement and adjusting
the action of the hydraulic system accordingly, draft-sensing systems
sense the pull force and determine the amount of load transfer. If the
pull becomes greater than a set level the system considers this as a
sign of the implement working too shallow and lowers the implement.
On soils with reasonably uniform specific resistance this system also
provides an even working depth for an uneven field surface, permitting
the implement to move up and down relative to the tractor in order to
maintain the constant working depth. If on the other hand, the specific
soil resistance varies in the field with some portions being soft and
others being hard, the draft-sensing system will give an uneven plowing
depth with a deeper working depth on the soft soil and a shallower
working depth on the hard soil.
The draft-sensing element in the three-point hitch is normally a spring
than can be located either in the top link (most often for small tractors)
or in the lower links (for the larger tractors). If the sensing element is
located in the lower links it can be used to maintain constant working
depth also for semimounted implement which normally do not have a
top link attached.

8.10 POSITION CONTROL

Another type of automatic implement depth control is designed such


that it maintains the position of the implement relative to the tractor
always at a fixed point. This point is normally set by the operator with
a stop for the control lever. When he moves the control lever down
against the stop the implement will move to the corresponding position
as soon as it can.
The position control system will maintain a fixed position of the
implement independent of the draft, in other words, independent of
the resistance of the soil. On the other hand, it will not compensate
8.12 Open- or Closed-Center Hydraulics 125

for an uneven soil surface, for instance, if the rear wheels go down in
a depression or if the rear wheels dig in due to high slippage.
Modern tractors are often equipped with combination of draft-sen-
sing and position-sensing capability such that the operator can decide
to what extent the implement position should be determined by draft
and or by the relative position. In this way he can find the combination
that gives the best result.

8.11 HYDRAULIC SYSTEMS

The control of mounted or semimounted implements requires such


forces that mechanical power is needed. Since the forces are high and
the displacements are often short, hydraulic power has been found to
be suitable because it can also be easily controlled and automated.
The hydraulic system consists basically of a pump, an actuator (cyl-
inder), valves, and piping.
The amount of power available from a hydraulic system can be cal-
culated from the pressure and the flow rate according to the equation
tv = TJPQ (8-8)

where W is the mechanical power (kW) from the cylinder or hydraulic


motor, TJ the cylinder or motor efficiency, P the available pressure
difference (MPa) at the cylinder and Q the available flow rate (1 sec - I)
at the cylinder.
The pressures and the flow rates in Eq. (8-8) are the effective flow
rates and pressures. These are always lower than the rated pressure
and the rated flow rate found in the specifications for the tractors.

8.12 OPEN- OR CLOSED-CENTER HYDRAULICS

The hydraulic pump and the pressure control valve in the system
have to be designed such that sufficient power is always available for
the operation of the implement when needed. When power is not
needed for the operation of the implement, the pump still delivers the
same amount of oil per unit time, the same flow rate. If the control
valve to the cylinder is closed the pressure in the rest of the system
will automatically go up to the value set by the relief valve. This would
126 8. Connecting Tractor and Implement

require considerable power from the engine and produce rapid heating
of the hydraulic oil.
The most common way of alleviating this problem is to use an open-
center control valve in the system. At the same time as the lines to
the cylinder are closed, a passage is opened for the oil from the pump
back to the tank with very little back pressure. The open center system
is a simple system and is commonly used if only one or two hydraulic
cylinders are used in the entire hydraulic system.
A common trend in modern practice has been to use the hydraulic
power for a number of functions other than raising and lowering a single
implement. Hydraulic power has been used for steering, for shifting
the transmission, for braking, and for operating remote cylinders from
the tractor. An open-center system may be used also for such cases
but has a number of operational disadvantages. Such systems are there-
fore often built as closed-center systems. In a closed-center system the
oil is not permitted to return directly to the reservoir at idling. Pressure
is maintained in the system and power loss and overheating of the
system is prevented by reducing the flow from the pump. The pump
is normally a variable-displacement pump, which is more complicated
than the fixed-displacement pump used in an open-center system.
A modern trend is to use electrohydraulic control valves for the oil
flow to the various cylinders and motors. These valves can be located
in the shortest path from pump to actuator. The oil does not have to
be taken up to the operator's platform, thereby simplifying the system
and reducing losses.
9
Operating the Tractor and
the Implement*

The human operator is still essential for an efficient use of the tractor
with attached implements. Its not likely that the operator will be re-
placed by mechanical devices in the near future even though mechan-
ical devices will gradually take over a number of the functions currently
performed by the human operator.
The tractor operator was largely ignored in the design of agricultural
tractors until around 30 years ago, when the first steps were taken in
improving the safety of the tractor operator and the working environ-
ment around him. It was soon found that the effectiveness of the entire
unit increased with a better workplace, better control, better environ-
ment, and safer operating conditions.

9.1 THE TRACTOR AS A WORKPLACE

Studies to make the workplace more efficient have been made in


other industries for a long time. One of the factors that has to be rec-

* This chapter was written by Sverker Persson, Agricultural Engineering Department,


Pennsylvania State University, University Park, PA.

127
128 9. Operating the Tractor and the Implement

Fig.9.1. Regions of operator visibility.

ognized in studying the tractor as a workplace is the matter of visibility.


Three areas of operator visibility can be defined and are illustrated in
Fig. 9.1. In area I, which is the principal area of attention, the operator
can monitor what is going on without moving his head and with full
use of his vision. Located in this area should be those functions that
require constant supervision and immediate attention from the oper-
ator. It also appears to be very important that the field of vision is not
interrupted by obstacles so that a reference line, for instance, a row
of plants, does not disappear behind a structural member of the tractor
and then appear again closer to the operator.
Since the main purpose of the operator on the tractor is to observe
what is going on and take action accordingly, his visibility must be
made as good as possible. Such obstacles as muffler, air cleaner, cab
frame, or implement frame should perferably be removed from this
primary area of vision or otherwise be made as small as possible.
The second area (II) with regard to visibility is an area that the op-
erator has within his normal 1800 field of vision, but outside the middle
area of primary visibility. The operator is aware of what is going on
in this area but has a less complete apprehension of observations in
this field. In this area could be located phenomena that require only
occasional action from the operator but of which he should be con-
stantly aware. Control handles, which often are located in this area of
9.1 The Tractor as a Workplace 129

Rear
hitchpoinls

~~~~~~~ Front hitchpoinls

Fig. 9.2. Regions hidden from the operator.

secondary vision, should be made such that the operator can recognize
them positively without looking at them, for instance, by giving the
handles different shapes.
The third area of visibility (III) is behind the operator. To take action
due to occurrences in this area requires the operator to turn around
and remove his attention from the primary area of observation. Ideally,
therefore, only such control functions that do not require operator ob-
servation should be located in this area.
Another area of concern, with regard to the operator visibility, is
that some areas around the tractor are hidden from his normal view.
Figure 9.2 shows an example of such an area. This diagram was created
by placing a light bulb in the normal position of the operators eyes and
tracing the shadows cast by the components of the tractor. If any vital
point is in the shaded area, activities at this point become much more
difficult than if it was fully visible. Points in this area of special concern
could be hitch points for connecting the tractor to implements. The
hitch point for heavy wagons is often in a hidden area. If the front
130 9. Operating the Tractor and the Implement

wheels are in the shaded (hidden) area, the precise steering of the
tractor is made more difficult.

9.2 TRACTOR CONTROLS AND INSTRUMENTS

Modern tractors are equipped with a large number of controls for


the operator to handle and also a large number of instruments to help
monitor what is happening on the tractor and the implement. The ques-
tion of efficient design of controls and instruments has earlier been
given great emphasis by the designers of military aircraft and other
military vehicles. The experiences of these studies can very well be
applied to agricultural tractors.
A general principle for the actuation of controls is that the most
frequently used controls should be actuated with a minimum movement
of the hand or the foot, while controls that are used very seldom can
require more movement of the arm, leg, or body. Consequently, con-
trols used very frequently should be located such that they can be
moved using only the finger joints, the wrist, the elbow, or the ankle.
Controls requiring movement at the shoulder, the knee, or the hip
should have to be used more seldom, while the least frequently used
controls could require movement of the entire body. This means that
the first group of controls should be located basically within a sphere
with its center at the wrist or elbow, the second group within a sphere
with a center at the shoulder and the third group further out to the limit
of reach (Fig. 9.3).
The above discussion on the different fields of visibility should also
be considered with regard to the location of instruments. The required
motions for both controls and instruments should be in a direction that
produces a consistent and expected effect. It appears logical that a
clockwise motion and an upward motion would indicate an increase in
the magnitude of the function to be controlled or monitored. Specific
guidelines are given in ASAE Standard S335.2 (ASAE, 1983a).
For some functions a warning light will give all the necessary infor-
mation, while in other cases a moving pointer or similar symbol will
give the operator better information about the parameter he should be
informed about. A moving pointer indicates more clearly how close a
variable is to its critical value than a warning light.
Some controls require a fair amount of muscle power in order to be
moved, e.g., a brake pedal. The direction of motion should be such
9.2 Tractor Controls and Instruments 131

mm
Shoulder pivot Far high 500
..... .....
--- 400
.-
\
-, Near
high 300
\
\ 200
Elbow Near
pivot low
\ 100
I Seat reference
I point
.0:-=-
=-
-
Reference line
----
I
- - - -
I
I
I)l,
'\,
mm
100
/
I
I

/
h
200

300

Fig. 9.3. Possible movements for control actuation. - , Heel; ---, toe.

that it is easy for the arm or the foot to move the control, which is
normally in the direction toward or away from the shoulder or away
from the hip. Figure 9.4 gives examples of good and bad control move-
ments from this standpoint.
If on the other hand the manual movement of a control requires a
fairly long movement, for instance, in making several turns of the steer-
ing wheel, the motion should be a swinging motion around the bodily
pivot points, which may be elbows or shoulders. Figure 9.5 gives an
example of a good and bad motion of a steering wheel that must be
moved through turns.
With the continuing growth in the size of implements and tractors,
many of the control functions that could earlier be made with only
human muscle power, today would require very great efforts from the
operator. It is thus common to install servo power. The tendency is in
the direction of controlling a machine with less and less physical effort.
Power steering has for a long time been common on tractors, and power
brakes, clutches, etc. are gradually being accepted.
The earliest need for improved control power came with the control
132 9. Operating the Tractor and the Implement

Good Bad

Good Bad

Fig. 9.4. Desirable and undesirable control movements.

of the implement from the tractor. As the implements became heavier,


it became impossible to raise and lower them manually even with good
linkage designs. Power control of implements is today practically stan-
dard on all tractors. Another reason for these functions is that the
distance between the operator and the implement has gradually in-
creased, making it very difficult to achieve control with manual de-
vices. The power for the control is often hydraulic but increasingly
electrical power is used for a number of implement control functions.
Due to the need for interchangeability between tractors and imple-
ments, a standardization of these remote-control devices has become
necessary and is to some extent available (ASAE 1983b).
Many controlled functions in the tractor are predictable and are re-
peated constantly, which makes them suitable for automation, meaning
that the control function is taken over by a mechanical device. Prac-
tically from the beginning of engine use on tractors the control of engine
9.2 Tractor Controls and Instruments 133

Easy

Difficult

Fig. 9.5. Desirable and undesirable steering wheel movements.

speed was given to an automatic governor. With the advent of the fully
mounted plow the automatic depth control for the plow was introduced.
Several attempts have been made to introduce automatic steering of
tractors but these attempts have not yet resulted in a commercially
acceptable solution. It is, however, likely that sooner or later automatic
guiding devices will be available. Other possible automatic control de-
vices that are being considered are devices that will choose the gear
and adjust the engine speed according to the power need of the im-
plement in such a way that minimum fuel consumption and engine wear
is achieved.
In general an automatic control device should be used (1) on func-
tions where the setting has to be adjusted continuously, (2) where a
suitable monitor is available for determining when change of setting is
needed, (3) where a simple relationship can be established between the
input from the sensor and the necessary control function. Automation
is also suitable where reaction by the human to an input is not suffi-
134 9. Operating the Tractor and the Implement

ciently fast or where the human is not capable of reacting in the time
required to detect that an adjustment is necessary.
Automation is less suitable and is difficult to apply for situations that
seldom arise and where there is no simple indicator of an adjustment
need, or where the decision about the adjustment has to be the result
of a comparison of several factors.
The aim of the automation should be that the operator is relieved of
all routine functions and given time to observe the general function of
tractor and implement. Modern computer technology will very likely
make it possible to automate many more functions in the operation of
a tractor.

9.3 OPERATOR ENVIRONMENT

The environment around the operator has great effect on his ability
to perform his normal functions efficiently. Among the environmental
factors to be considered are vibration, shape of the seat, noise, tem-
perature, and air quality.
Tractor operation has for a long time been connected with severe
vibration problems caused by driving an unsprung tractor over uneven
fields. The human body has the ability to withstand considerable vi-
brations and has its own means of protecting vital organs from being
damaged from such outside vibrations. It was, for a long time, con-
sidered desirable to make the operator's platform such that the operator
could drive in a standing position thereby using his legs to reduce some
of the vibrations from the tractor. With the more crowded conditions
in today's operator stations, this possibility is normally eliminated.
The vibrations affect the operator both with regard to their amplitude
and frequency. A common measure for describing the severity of the
vibrations is to indicate the accelerations. Figure 9.6 shows the sen-
sitivity of the body to vibrations of different accelerations as a function
of frequency. It should be noticed that the sensitivity is very high at
a frequency around 5 Hz. This indicates that the ability of the human
body to resist the vibrations is fairly low at this frequency. It is un-
fortunate that a frequency of about this magnitude is fairly common
on tractors, because tractors are supported by high-volume, low-pres-
sure rear tires. Modern tractor seats greatly reduce the problems of
vibrations. Such seats have a sophisticated suspension system, in-
cluding springs and damping devices, and both of these are adjustable
9.3 Operator Environment 135

1.0 , - - - - r- - r -- - - r - , - - - , - - , - - -r---r--.Y7''T7777r-,
0.8 f-------+---+--+--+-++---t--+--r.l~nlS-5M---1

0.6 f-------+---+--+--+-++---t-.....+~~;g.~--1
'"

0.05 f---+-+--I-~~I-I---I--t-+----f---1H

2346810 20 30 40 60 80 100
Frequency, Hz

Fig. 9.6. Operator sensitivity to vibrations.

according to the weight of the tractor operator. Modern tractor seats


also have been given a shape that will support the operator's body in
such a way that he gets less tired.
Another factor of concern is the noise level around the tractor op-
erator. Modern requirements set a noise level of 85 dB as the maximum
level for 8 hr daily exposure. If the noise level goes to higher values,
the length of exposure time should be reduced according to the fol-
lowing:
hours dB

8 85
4 90
2 95
100

Improving the environment around the tractor operator also includes


providing suitable temperature by either air conditioning in the summer
or heating in the winter. Further, protection against dust is needed
under some operating conditions.
Naturally the environment should not include any poisonous or nox-
ious gases for instance, fumes from gear boxes or exhaust from the
engine. Certain guidelines for the placement of exhaust pipes and gear-
box vents have to be considered with regard to these factors.
136 9. Operating the Tractor and the Implement

9.4 OPERATOR SAFETY

Statistics of work-related accidents show that agriculture is at the


top, or very close to the top, with regard to work-related accidents per
worker both for fatal accidents and serious, nonfatal accidents. Since
tractor operation is one of the major activities on the farm, it is natural
that the tractor is involved in a substantial number of these accidents.
Up until recently, approximately two-thirds of the serious tractor ac-
cidents involved the tractor overturning either sideways or backward.
Other major accident categories involve the operator or some other
person falling off the tractor and being run over, and injuries resulting
from being caught in rotating shafts.
Due to the desire to concentrate as much load as possible on the
drive wheels of a tractor and due to the uneven and unprepared terrain
over which a tractor operates, the tractor is basically an unstable ve-
hicle and can overturn more easily than many other vehicles. (This
was discussed in more detail in Chapter 7.) Even with normal caution
a tractor operator may sometimes get into a situation that is dangerous.
It is often not a single factor that makes the tractor overturn but a
combination of factors, for instance, driving on a slope and hitting a
depression in the ground, starting the tractor out of a ditch, or carrying
a large bale in the front-end loader and hitting a stump or mound in
the field.
Tractor operation is essentially a one-man operation that often oc-
curs far away from other workers; consequently a tractor operator
involved in an accident may not get help until it is too late.
Many devices have been suggested for preventing the tractor from
overturning. However, arrangements that increase the basic stability
of the tractor may prevent it from performing normal functions. De-
vices that disconnect the power before a dangerous situation has de-
veloped have not been practical either. The time available for pre-
venting a dangerous situation from becoming an accident is normally
too short for human reaction.
Consequently it has been found that the best way to avoid injury and
damage to the operator is to provide him with protection in case a
tractor overturns. Such rollover protection devices are presently re-
quired by law in many countries and are highly recommended in other
countries. Requirements have been established for testing such devices
to make sure that they have sufficient strength. A rollover protection
9.5 Systems for Hitching the Implement to the Tractor 137

device that is not sufficiently strong is worse than no protection at all


because it gives the operator a false sense of security.
The principal purpose of the structure is to be able to absorb the
anticipated impact energy that may result from overturning without
deforming to such an extent that the operator gets hurt. Earlier testing
procedures for rollover protection structures were built on impact test-
ing by providing a similar kind of load to that expected in an overturn
situation. Some of these tests have now been replaced by other more
easily controlled static tests.
Most tractor manufacturers have found it economically favorable to
combine the rollover protection structure with at least a roof or a wind-
shield and sometimes a fully enclosed cab with air conditioning.
The effect of the introduction of rollover protection devices has been
a dramatic reduction in the number of fatal and serious accidents as-
sociated with tractor operations.
A number of accidents happen when the operator is hitching an im-
plement to the tractor or getting on and off the tractor in order to
connect an implement. Improved hitching systems, such as those out-
lined in the next section, are needed in order to reduce such accidents.

9.S SYSTEMS FOR HITCHING THE IMPLEMENT


TO THE TRACTOR

Most tractor hitches are designed with regard to the proper transfer
of forces and motions between the tractor and the implement during
work. Lately the question of how to connect the tractor to an implement
has also been given consideration.
Normally the implement to which a tractor is to be connected is not
lined up well enough with the tractor at the time of hitching to make
it possible for a direct link-up. This means that the position of the
implement has to be adjusted relative to the tractor. With earlier, fairly
light implements this could reasonably be done manually by the op-
erator. However, with modern, much bigger implements, it may be
beyond the ability of the operator to line up the implement behind the
tractor and other means are required to bring about the hitching. A
goal should also be to enable hitching to be carried out without the
operator leaving his seat.
Among the principles applied in the design of improved hitching sys-
tem should be mentioned the following:
138 9. Operating the Tractor and the Implement

--_ _ _-3a

3b

Fig. 9.7. Driving catch space.

1. Points to be connected by the operator must be visible from the


tractor driver's seat. They must also be located such that the driver
can easily maneuver the tractor to them. This requirement can be il-
lustrated by the concept of the driving catch space as shown in Fig.
9.7. In tractor position a, the driving catch space for hitchpoint 2a is
limited by the line 1, 2a, 3a, and includes both the implement points
A and B. In the subsequent tractor position b, point 2a has moved to
2b, the driving catch-space has moved to 1, 2b, 3b and no longer con-
tains implement point B. Implement point B can then not be connected.
It should be noticed that if the hitch point is in the rear end of the
tractor, the tractor driver has very little opportunity to move the hitch
sideways when he gets close to the implement.
2. A second possibility for making the hitching easier is to have
the hitch point powered such that it can be moved to the proper location
using hydraulic or other power with the operator still in the driver's
seat. Examples of such are the lower links in the three-point hitch,
which can be moved up and down using hydraulic power. On some
tractors these lower hitch points can also be extended by the operator
before he drives the tractor up to the implement and then pushed into
working position after they are connected by the operator driving the
9.5 Systems for Hitching the Implement to the Tractor 139

Fig. 9.8. Possible misalignments during hitching.


140 9. Operating the Tractor and the Implement

tractor back against the implement. Such powered motions of the hitch
points are desirable in order to simplify hitching. For example, powered
motion sideways of the two lower hitch points on the tractor and also
the possibility of powered extending or shortening of the top link should
be considered.
3. Make the contact points self-guiding such that they would au-
tomatically adjusted relative to each other. A common example is the
funnel used for the hitch hook on many trailers.
When designing hitches for better hitching, it is necessary to have
some dimensions for likely misalignments between tractor and imple-
ment during normal hookup (Fig. 9.8).
Eventually some kind of universal standard will likely be accepted
that will list to what extent the hitch design should accommodate mis-
alignment between tractor and implement.

REFERENCES
ASEA (1983a). Operator controls on agricultural equipment. ASAE Standard S335.2, in
"American Society of Agricultural Engineers Yearbook." ASAE, St. Joseph, Ml.
ASAE (1983b). Application of hydraulic remote control cylinders to agricultural tractors
and teaching type agricultural implements. ASAE Standard S201A, in "American Soci-
ety of Agricultural Engineers Yearbook." ASAE, St. Joseph, Ml.
10
Developments in the
"Tractor" Concept

10.1 INTRODUCTION

The conventional two-wheel drive tractor can be crudely described


as consisting of two large-diameter drive wheels that are rigidly at-
tached to the transmission housing and frame supporting the engine.
The implements are attached behind the rear wheels, and for mounted
implements the engine acts as a counterweight, enabling the implement
to be lifted. The tractor is intended to be versatile. It is used for a range
of field activities, for transport, and for chore tasks around the farm-
stead. The function of the tractor is to provide the power necessary
for various field and transport activities.
Power units used in industry are usually stationary or move through
a fixed directrix and limited range, e.g., the overhead crane. An im-
portant distinction between extensive forms of agriculture and the usual
industrial setting is that in industry materials are carried to stationary
machines. This contrasts with agriculture, where a universal machine,
i.e., the tractor, is typically brought to the materials. Hence the so-
lutions applicable to industry are often of little benefit to agricultural
mechanization.
Field operations required for crop production can be defined, ge-
141
142 10. Developments in the "Tractor" Concept

nerically, as the control of the implement relative to the crop. This


broad definition can be broken down into specific categories:
1. controlling and powering tillage implements usually via the draw-
bar,
2. controlling and powering field implements using the PTO as the
source of power,
3. transportation of materials to be collected or distributed from or
to an area.
In addition to these field functions, the tractor is also used in trans-
port mode. The transport functions can be categorized as:
4. road hauling, which mayor may not be an extension of field
hauling,
5. chore tasks around the farmstead, usually involving the use of
a loader.
This mUltiplicity of requirements necessitates comprises in the de-
sign of a tractor. The resulting limitations imposed in terms of the ability
of the tractor-implement system to provide the high-output capacity
required for some field activities has led to the adoption of some spe-
cial-purpose, self-propelled machines. These machines, such as the
combine harvester and the self-propelled forage harvester, have been
designed to meet the requirements of a specific task.
The basic traction requirements call for the ability of the vehicle to
develop maximum drawbar power with minimum soil compaction, roll-
ing resistance, and wheel slip. These requirements are best met by
having large-diameter, low-pressure tires; however, the specific re-
quirements for the tractor will significantly influence the final selection.
Traction is only one requirement and may indeed be of secondary con-
sideration relative to other requirements, such as speed of operation,
overall vehicle height, cab location, and loader attachment. Reece
(1969) developed a chart to show the inter-relationships between the
types of traction device and the major variables of soil type, duty re-
quirements, and speed of operation. These relationships are given in
Table 10.1
The compaction effects of operating with high-powered and heavier
tractors is of some concern. Compaction reduces the macropore space,
which provides the oxygen necessary for root development. Compac-
tion also impedes the infiltration of surface water, resulting in the in-
creased likelihood of runoff and surface erosion. The effects of com-
10.2 Controlled-Traffic Farming 143

Table 10.1. Relationship between Traction Device, Soil Conditions, Vehicle Duty,
and Speed (Reece, 1969)

Traction device Soil Duty Speed

Rigid wheel or high-pressure tire Hard, sandy Carrying Fast


Soft tire t t t
Track Soft clay Traction Slow

paction are likely to be more pronounced on wetter soils; however,


most agricultural soils will show some compaction effects in the form
of suppressed root growth and reduced yields. These effects can be
countered, in part, by using chisel plows and subsoilers to break up
compaction zones. The act of subsoiling can, of course, create con-
siderable smearing and localized compaction because of the heavy
drawbar loads imposed on the tractor. A suggested alternative strategy
is to use the Wye Double Digger (I. Warboys, personal communication,
1982), which consists of a plow and rotary cultivator. The cultivator is
power driven and is designed to operate in the plow bottom. The rotary
cultivator operates in the open furrow created by the plow's previous
pass. The net effect is a double-digging arrangement, which is claimed
to remove the plow pan, improve soil drainage, and increase crop
yields.

10.2 CONTROLLED-TRAFFIC FARMING

Controlled-traffic farming is typified by those machinery systems


which reduce the frequency of passage of tractors and machinery over
the cropping area. This can be accomplished by providing semiper-
manent traction lanes and cropping between them using a powered-
tool frame built as a gantry. The gantry concept is typified by that
proposed by Chamen et al. (1980), who proposed a wide-span gantry
to replace the field tractor. The gantry was seen as providing the po-
tential for operational control, efficient cultural practices, and reduced
implement costs. The gantry would use preestablished pathways in the
field and a guidance system to maintain direction and compensate for
differential slip between the driving wheels on either end of the gantry.
The operator's cab would be capable of traversing the width of the
144 10. Developments in the "Tractor"' Concept

gantry and pivoting through 3600 • The function of the operator would
be to monitor the performance of the implements.
It is interesting to note that the early concept of winch-plowing,
which involved the winching of a plow via a cable between two steam
engines, provided a method of cultivating without having to move the
power unit (the steam engines) across the field. A recent version of
the cable-winching concept has been adopted at Texas A&M Univer-
sity (Le Pori et al., 1983). This version of the cable-winching approach
consisted of a power unit that was located in position at one end of
the field. This unit powered the winch to pull the cable and attached
implement across the field. A mobile truss, similar to that used for
linear-move irrigation systems, was positioned along the length of the
cable to provide directional stability for the trailed implement. In the
steam engine winching systems, the side draft forces were countered
by the man walking behind and steering the winched plow. The ad-
vantages claimed for this system are that it lends itself to field auto-
mation, it provides a method of controlled traffic farming, and it elim-
inates the slip and rolling resistance losses associated with conventional
tractive systems. Another possible advantage is that electricity could
be used to supply the energy for the power unit.
The gantry concept is also demonstrated in the "cross-spanner" ma-
chine developed by PSI Engineering (W. L. Spratt, personal com-
munication, Panel Systems, Inc., Hurricane Mills, TN, 1983). This unit
consists of a mainframe, an overhead carriage, and a range of imple-
ments. The carriage is suspended from the overhead frame and carries
the implements. The carriage is self-propelled and is powered along
the mainframe, thereby causing the carriage and attached implement
to move over the ground. Having made a "pass" with the carriage,
the mainframe is indexed forward by a distance equal to the width of
the implement. It is envisaged that this system would have application
not only for cultivation but also for harvesting and some primary sorting
or processing. It should be pointed out that this machine is still in the
development and prototype testing phase.
An approach used by the author (Alcock and Jahns, 1984) was to
use the boom of a center-pivot irrigator as a position-referencing device
for an electrically powered tool frame. This was essentially a guidance
system based on the master-slave concept with the "master control"
provided by the boom of the center pivot. The "slave" was the elec-
trically powered tool frame, which was powered and steered by indi-
vidual wheel motors. As the boom moved through an arc, the tool frame
10.2 Controlled-Traffic Farming 145

Irrigator pipe

Control switches

Control panel

Control
switches

Output shaft
(to wheel)

Fig.10.1. Schematic of the tool frame and power supply.

followed in a spiral pattern. The polar radius was adjusted continuously


so that its value changed by an amount equal to the width of the im-
plement after each complete revolution. Figure 10.1 is a schematic of
the tool frame and power supply. The electrical supply to the tool frame
is carried through a metal-sheathed "umbilical cord," which also
serves as a distance control actuator to position the tool frame at the
correct distance from the boom. The orientation of the tool frame was
controlled by microswitches located at the end of the umbilical cord
on the tool frame.
146 10. Developments in the "Tractor" Concept

Transmitter -~..,

Microprocessor

Fig. 10.2. Spiral control of tractor and implement.

The system proposed may have some unique applications, for ex-
ample, in the turf-growing industry, where repetitive mowing is es-
sential and large center-pivot irrigators are commonly used. Gang mow-
ers could be operated on a continuous basis from the tool frame as it
follows the spiral path from the outer radius to the inner effective ra-
dius. It is also envisaged that tractor-implement combinations could
be controlled to follow a spiral pattern by using a centrally located
transmitter with an on-board receiver and microprocessor to check the
polar radius in accordance with the desired path, as illustrated in Fig.
10.2.

10.3 BATTERY-POWERED TRACTOR

The use of electricity as the energy source for agricultural vehicles


has received very little attention. Reece (1968) proposed an interesting
10.3 Battery-Powered Tractor 147

Table 10.2. Comparison of Fuel Energy Densities

Equivalent to
Energy density Conversion efficiency 100 liters of gasoline
Fuel source (MJ/kg) assumed (kg)

Petroleum 44.2 0.20 74.1


Diesel 43.0 0.26 58.5
Lead acid battery 0.108-0.144 0.85 5344-7125
Advanced lead-acid
battery 0.180-0.216 0.85 3563-4276

system comprised of permanent tracks around the fields in which would


run electrically powered cars equipped with electrically driven winches
for pulling the tillage equipment between cars and across the field. Allis
Chalmers built an experimental electric tractor powered by a fuel cell
(Thrig, 1960). It had a 15-kW electric motor and developed a drawbar
pull of approximately 13 kN.
The major difficulty associated with powering a tractor via energy
supplied from a battery is its low energy density. Table 10.2 compares
the energy density of some commonly used fuels based on their energy
equivalency to 100 liters of gasoline. It is readily apparent that, even
taking into account the higher conversion efficiency of the electric
drive, the battery mass requirement is considerable. There have been
some recent improvements, however, in battery technology both in
terms of the number of cycles to be expected and the available energy
density. Flat-plate traction batteries of the lead-acid type typically
provide an energy density of 24 Whr kg-I. Tubular plate lead-acid
traction batteries that provide an energy density of 38 Whr kg-I are
now available (Lucas, 1983). This represents a significant improvement
in the energy density and offers some real prospects for the commercial
application of battery-powered vehicles.
A battery-powered tractor developed at South Dakota State Uni-
versity (R. Alcock and L. L. Christianson, 1985) is intended for chore-
type tasks, which typically involve loader and some haulage work. The
battery-powered tractor does offer some unique advantages over a
comparable diesel-powered version. It is much quieter and is subject
to much less vibration, which improves the comfort of the operator.
It is also pollution free and can therefore be used within buildings and
near people and livestock without causing them any discomfort. The
148 10. Developments in the "Tractor" Concept

Table 10.3. Energy Use Comparison

Cost
Task Description Energy use (MJt (¢) % of diesel cost

Diesel-powered tractor
I Full load 4.79 4.06b
2 Part load 14.13 11.99b
3 Full load 16.50 14.00b
4 Part load 23.93 20.31 b
Battery-powered tractor
1 Full load 2.05 c 3.4l d 84.0
2 Part load 3.38 c 5.64d 47.0
3 Full load 6.75 c 11.25d 80.3
4 Part load 6.78C 11.3if 55.6

a Average energy use for five repetitions.


b Based on 30¢ per liter.
c Assuming a battery and charger efficiency of 70%.
d Based on 6¢ per kWhr.

maintenance costs are significantly less because of the simplicity of


the dc series-wound motor that is used to provide traction. A second,
smaller motor supplies the power for a hydraulics package and to drive
the PTO shaft. Table 10.3 provides a comparison of the costs associated
with diesel and battery tractors of similar configuration and size per-
forming the same tasks. The battery-powered tractor has an energy
consumption that is significantly less than that of the diesel tractor,
particularly when operating under part-load condition.

10.3.1 Batteries
The battery pack used in the battery-powered "choremaster" tractor
consists of two 32-cell blocks, providing a nominal operating voltage
of 128 V. They provide a total battery capacity of 340 A hr at the 6-
hr rate, or 43.5 kWhr. Each battery block is 0.89 m in length, 0.5 m
in width, and 0.59 m in depth. The total battery mass is 1914 kg, with
an effective energy density of 24 Whr/kg. This energy density is low
when compared with the advanced lead-acid batteries developed by
Lucas Chloride, which have an effective energy density of 38 Whr/kg.

10.3.2 Traction Motor


The traction motor was wound to provide the torque characteristics
required. The tractive effort vs. speed curves for the vehicle are given
10.3 Battery-Powered Tractor 149

100

90
First gear

80

70

Z 60
~
1:: Second gear
0
==OJ 50
OJ
>
U
ro 40
F

30 Third gear

20

10

0
5 10 15 20 25
Vehicle speed (km/hr)

Fig. 10.3. Tractive effort vs. vehicle speed.

in Fig. 10.3 and the characteristics of the motor are shown in Fig. 10.4.
The operating speed range of the motor is between 930 and 2750 rpm.
Operating below 930 rpm, at nominal motor voltage and at full load,
is precluded by the large current demand. Operation at speeds above
2750 rpm is prevented by an "overspeed" switch. A speed control unit
mounted in the tractor cab can be used, however, to provide infinite
speed variation from 0 to 2750 rpm. The traction motor has a 1-hr rating
of 37 kW, and can also supply 81 kW for 3.3 min. Thermistors in the
motor frames are used to control the operation of small, 12-V blowers
for cooling both the traction and PTO/hydraulics motors. The PTO/
hydraulics motor is rated at 17 kW for 1 hr and has overload charac-
teristics similar to the traction motor. In the present vehicle, this motor
drives a hydraulic pump for steering and hydraulics, as well as the PTO
shaft. The hydraulic pump is fitted with a priority valve to ensure oil
supply to the steering rams.
150 10. Developments in the "Tractor" Concept

100 1000

80 800
~
>-
u
c:
Q>
'0 60 600 E
!E I
Q> ~
'C Q>
c: :::J
'" e-
40 400 F2
~
~
~
20 Power 200

Torque

0 0
500 1000 1500 2000 2500 3000
Motor speed (rpm)

Fig. 10.4. Electric motor characteristics.

10.3.3 Controllers
Speed control of the traction and PTO motors is provided by two
silicon-controlled rectifier (SCR) controllers with contacters and a by-
pass circuit for full-speed operation. The controller for the traction
motor can provide 265 A continuously (producing 34 kW) and 550 A
for 5 min (providing 70 kW). The controller for the PTO/hydraulics
motor gives 250 A continuously (26 kW) and 475 A for 5 min (61 kW).
Two lever-actuated potentiometers positioned in the tractor cab enable
the operator to adjust the speeds of each motor. A reversing switch,
operated from a cam on the speed control lever, allows the operator
to change the direction of the traction motor for quick and easy re-
versing of the tractor. The reverse mode can also be used to provide
a braking effect against the direction of travel. The motors used are
sensitive to change in load, resulting in a reduction in motor speed as
the load is increased. This is evident from Fig. 10.4. A feedback circuit
that will allow a set motor speed to be maintained has been installed.
This is accomplished by varying the voltage at the motor in accordance
with change in output speed resulting from an applied load. In reality,
10.3 Battery-Powered Tractor 151

Table 10.4. Tractor Transmission Ratios

Gear Overall ratio Speed range (km/hr)

1 72: 1 0-8
2 36.4: 1 0-16
3 17.67: 1 0-24

the voltage applied to the motor terminals is full battery voltage. How-
ever, the switching provided by the SCRs varies the mean voltage
supplied as input to the motors.

10.3.4 Transmission
The output from the traction motor feeds into a three-speed transfer
box and then to the front and rear differentials. Power is transmitted
from the differentials through drop boxes mounted at each wheel. The
overall gear ratios and associated speed ranges are given in Table 10.4.

10.3.5 Battery Management


Instrumentation provided in the cab displays the state of key com-
ponents. A voltmeter registers the battery voltage at the controller and
ammeters display current draw for both the traction and PTO/hydraulic
motors. Also included are tachometers for each motor. An important
item of instrumentation is the "fuel gauge," which gives the operator
a useful indication of the state of charge of the battery. Initial battery
condition is used as a reference point and from this value the watt-
hour consumption is subtracted. When the battery has discharged to
20% of its rated capacity, an indicator light flashes on the fuel gauge.
The battery condition is also checked from the electrolyte specific grav-
ity readings. Specific gravity of the electrolyte in all of the cells is
measured once per month. In addition the specific gravity in two pilot
cells, one in each battery block, is checked on a daily basis. The battery
is recharged when the battery capacity has been reduced to 20% of its
nominal rating. Recharging takes from 6 to 8 hr, depending on the final
discharge state of the battery.
152 10. Developments in the "Tractor" Concept

10.3.6 Summary
Battery-powered vehicles are suited to intermittent operation, per-
forming tasks whose duty requirements are fairly readily identifiable.
The battery-powered, chore-type tractor described here may well find
application in materials-handling tasks for the logging, mining, food,
forestry, and building industries, as well as for certain materials-han-
dling tasks in the farming industries. The battery-powered tractor is
unlikely to be suited to extensive or continuous work such as field
tillage or road haulage over considerable distances. Its applications are
somewhat unique, and such a vehicle may complement, rather than
replace, the larger tractors intended for tillage work. The character-
istics of dc series-wound electric motors are such that they provide a
rising torque characteristic with fall in speed (due to an applied load).
This effect can be used to overcome temporary overloads, and when
used in conjunction with a suitable transmission, can provide a constant
power output at the drive wheels (or PTO).

10.4 SYSTEMS TRACTOR

Although the conventional tractor design seems unlikely to change


radically in the next two decades, there is some evidence that the so-
called systems tractor will increase in use, especially as the larger man-
ufacturers of tractors develop designs based on this concept. Systems
tractors are characterized by designs that enable field operation in
either direction, are equipped with front and rear hitching systems, as
well as PTO outlets at the front and rear of the tractor (Gohlick, 1984).
This type of tractor will be more suited to both field as well as transport
tasks than the conventional tractor. Front- and rear-mounted imple-
ments can be used simultaneously for field work; loader attachments
facilitate materials movement around the farmstead, and the cab lo-
cation and visibility permit equally good operation in both directions.
The systems tractor might also be used for high-speed road work, at
speeds up to 40 km hr- 1. This would allow the systems tractor to be
used effectively and economically for road transport, but would require
a totally suspended cab and the ability to switch from four-wheel drive
(for field work) to two-wheel drive (for road work).
References 153

REFERENCES
Alcock, R., and Jahns, O. (1984). Field automation by spiral mechanization. ASAE Paper
No. 84-1030. Am. Soc. Agric. Eng., St. Joseph, MI.
Chamen, W. T. C., Collins, T. S., Hoxey, R. P., and Knight, A. C. (1980). Mechanization
opportunities likely to be provided by engineering in the 21st century. The Agricultural
Engineer 35(3), 63-70.
Oohlich, H. (1984). The development of tractors and other agricultural vehicles. J. Agric.
Eng. Res. 29, 3-16.
Le Pori, W. A., Mizrach, A., Harrison, C. R., Evans, M. A., and Carney, D. B. (1983).
Functional evaluation of field machinery using a cable towing system. ASAE Paper
No. 83-1545. Am. Soc. Agric. Eng., St. Joseph, MI.
Lucas (1983). Lucas Pub. 4135, Lucas Chloride E. V. Systems Ltd, Birmingham, UK.
Reece, A. R. (1968). An automatic electric arable farm. Farm Mech. Buildings 20(226),
18-20.
Reece, A. R. (1969). The shape of the farm tractor. Proc. Inst. Mech. Engr. 184, part
3Q,125-131.
Thrig, H. K. (1960). An electric-powered tractor. Agric. Eng. 4, 232-233,240.
Alcock, R., and Christianson, L. L. (1985). A battery powered tractor Agric. Eng. 40(2),
69-73.
Appendix: Problems and
Problem-Solving

It is recommended that problem solutions should be developed, as far


as is practicable, using computer programs. Problems for solution
should therefore be of such a nature that they can be structured for
computer programming.
There are two possible approaches. The first is to start with a simple
problem that can be developed in complexity as solutions are devel-
oped. This allows confidence in the use of solution solving techniques
to be developed. The second is to offer a problem that is more complex
and is best tackled by initially simplifying or making simplifying as-
sumptions that permit the initial programs to be structured. This initial
program would then be added to until an acceptable representation of
the problem had been achieved. All the program examples given are
incomplete in that they require, in most cases, the specification of the
inputs necessary to enable the relevant calculations to be made. The
intention is to encourage the individual development of the necessary
program content, rather than provide a simple format.

Example 1.1
Develop an interactive computer program that allows data from dy-
namometer tests to be input and calculates power, torque, specific fuel
154
Appendix: Problems and Problem-Solving 155

consumption, and thermal efficiency. Provide an output that lists the


input data and calculated parameters in a tabular format.

Example 1.2
Add to the above program so that a graphic output, with power,
torque, specific fuel consumption and thermal efficiency, plotted on
the ordinate, and shaft speed on the abscissa, is produced as each set
of input data is entered.

Example 1.3
Add to the above by providing outputs giving the mean and standard
deviation of each variable, the rated power and speed, and the speed
at which maximum output torque occurs.

Example 2.1
Write an interactive program that will determine the gear ratios nec-
essary for a tractor to achieve a specified drawbar pull at rated engine
speed and to achieve a specified maximum forward speed.

Example 2.2
Modify the above program to allow for other inputs that will influence
the drawbar pull capability: the operating grade, the coefficient of roll-
ing resistance, and the maximum vehicle acceleration.

Example 2.3
Assuming that the gear ratios are spaced according to the geometric
progression rule, modify the program for Example 2.2 additionally to
determine the appropriate overall gear ratios for a given number of
gears. For each gear ratio, plot tractive effort (ordinate) vs. engine
speed (abscissa).

Example 2.4
Add to the above program so that the wheelslip associated with a
given tractor weight and thrust requirement can be determined. Use
the Wismer and Luth prediction equations and provide for an input of
the soil cone index value. Determine the wheelslip value. Determine
156 Appendix: Problems and Problem-Solving

the wheelslip values for the tractive effort determinations made in Ex-
ample 2.3.

Example 2.5
Add to the above program by providing the input capabilities nec-
essary to determine the tractive efficiency. Provide an output format
that will plot tractive efficiency as ordinate vs. wheelslip as abscissa
for different values of soil cone index.

Example 2.6
Repeat subprograms 2.4 and 2.5 using the Gee-Clough equations.
Compare these results with those obtained previously.

Example 2.7
Write an interactive program that will calculate the load inertia re-
flected to the engine output shaft. (Ignore the inertia load effect caused
by the transmission.) Calculate the reflected inertia for each of the
tractor overall gear ratios, at the engine speed corresponding to max-
imum torque, and determine the tractor accelerations expected and the
engine output power requirement.

Example 3.1
Write a program that determines the costs associated with machinery
ownership (fixed costs), time costs, and task costs. This program
should then be extended to determine the optimum machine capacity.

Example 3.2
Modify the above program to determine the optimum machine ca-
pacity for the analysis conducted in Example 3.1 if two machines of
similar capacity are purchased instead of one.

Example 3.3
Write a program to compare group replacement with individual re-
placement of items, such as cultivator tines or sprayer nozzle tips.
Determine if and when group replacement would become a viable al-
ternative to individual replacement. Assume that the failure rate fol-
lows a negative exponential distribution.
Appendix: Problems and Problem-Solving 157

Example 3.4
A battery-powered vehicle has a total weight of 50 kN and a battery
mass fraction of 0.4. The available energy density of the battery is 30
Whr/kg. The vehicle is four-wheel drive with equal-sized wheels of
1.10-m effective diameter and 0.35-m width. The soil cone index is 700
kN/m 2 • In each cycle, the vehicle is accelerated from rest on level
ground for 3 sec, then held at this velocity for 10 min, and then brought
to rest. A polynomial equal relating torque (M), in Nm, of the electric
drive motor and its output rpm (N) is as follows:
N = 0.000041M3 + 0.048M2 - 18.52M + 3412.3
The overall reduction ratio is 72: 1. The inertia of the motion is 0.173
kg m2 and the inertia of the wheels is 6.0 kg m2 • The vehicle has a
hydraulic system, for steering purposes, that requires 8 kW and has
an efficiency of 0.65. The transmission efficiency is 0.8, and the battery
and controller efficiency is 0.9. Write a program to provide perform-
ance data for the vehicle, including the expected number of cycles.
Index

A D
Acceleration, 98 Depreciation, 75
torque required for, 98 Dimensional analysis, 53
Adiabatic index, 28 Draft,41
Adiabatic process, 20 Drawbar
Aerodynamic resistance, 96-97 performance, 92
Air, constituents of, 18 pull, 36, 49, 52
Air flow, measurement of, 34 Dynamics, 6
Dynamometer, 35

B E
Ballast, 59, 61-67, 70-72 Efficiency
number, 65 mechanical, 24
Bathtub curve, 80 thermal, 15
Bekker theory, 49 tractive, 25, 54, 55, 57
Boyle's law, 16,20 transmission, 24
Breakdown rate, 77 Energy, 6
balance, 30
conservation of, 8
flow, 11
C kinetic, 8
Calorimeter, 32 nonflow equation, 13
Camot cycle, 23, 28 potential, 8
Charles' law, 16 steady-flow equation, 10
Cone index, 53 used in agriculture, 4
Cone penetrometer test, 53 Engine
Controlled traffic, 143-146 heat balance, 32
Coulomb equation, 46-48 thermodynamic, 15

159
160 Index

Enthalpy, 12 Hydraulic system, 125, 126


Entropy, 15,23 draft sensing, 123, 124
Ergonomics, 127-140; see also Operation lift capacity, 123
open and closed center, 125, 127
position control, 124, 125
F
Failure
probability, series and parallel I
arrangements, 82, 83 Implement, I, 37
rate, 79 control, 105, 106
Field Inertia, of the vehicle, 98, 99
capacity, 83 reflected, 98
optimum, 84, 86 resistance, 94-96
Fixed costs, 73, 74-76 Inflation rate, 74
Four stroke, 28 Insurance costs, 75
Fuel Interest rate, 74
calorific values, 31 Investment credit, 74
consumption, 76
specific, 33
energy density of, 147 L
flow, measurement of, 33 Load
distribution, 117-120
transfer, 113-120
G limits of, 116-118
Gantry, 4
systems, 143, 144 M
Gas Machinery
constant, 17 management, 73-91
laws, 16 selection, 83-88
processes, 19 Maintenance
constant pressure, 19 costs, 76
constant temperature, 20 frequency of, 77, 78
constant volume, 19 Mean effective pressure, 27
Governed range, 26 Micklethwaite equation, 46
Guidance systems, 144-146 Mobility number, 56, 57, 71
spiral pattern, 144-146 Mole, 17
Morse test, 29
H
Hitch N
functions of, 109-11 0 Nebraska test report, 38-41
mounted implements, 111 No-load speed, 26
point, 111, 112
virtual, 112
semimounted implements, 110 o
systems, 109-126, 137-140 Operating costs, 76-78
Housing costs, 75 Operation, of tractor and implement,
Hunt model, 84 127-140
Index 161

Operator Sinkage, 47, 52


controls and instruments, 130, 131 Slip, 51, 52
environment, 134-136 Soil deformation, 50
movements, 130-134 Specific heat, 18
safety, 136, 137 constant pressure, 19
visibility, 127-130 constant volume, 19
Stability, longitudinal, 99
stabilizing factor, 100
p Swept volume, 28
Plastic limit, 46 Systems, 5
Power System boundary, 6
brake, 29, 24 thermodynamic, 11
drawbar, 35, 92, 93; see also Drawbar
performance
friction, 29 T
indicated, 24, 27, 29 Thermodynamics, 10
measurement, 24 first law of, 14
second law of, 14
PTO, 3, 24
Three-point hitch, 120-123
Power-to-mass ratio, 4
Thrust, 51-53
Pressure-volume diagram, 21
Tires, 59, 60, 68
Problems, 154-157
selection of, 68-72
PTO, 35; see also Power, PTO
Torque, 25, 26, 41, 53, 97
Pull, 53, 105
back-up, 26, 27
drawbar,92
Towed force, 53
maximum, 55, 56
Traction, 44
ratio, 57 coefficient of, 57-66
Tractive effort, 36, 93, 149
R Tractor, 1, 3
Rated engine speed, 26, 27 battery powered, 148-151
Reliability, 78-83 and compaction, 142, 143
Repair times, 77 developments in, 141-152
Replacement theory, 86-91 dynamics, 101-103
challenger, 87 functions of, 142
defender, 87 performance, 35
group, 88-91 systems, 152
holding cost, 87, 88 Tractor-implement, 2
technological impacts, 88 forces in, 103-108
Resale value, 75 Transfers, energy and information, 7
Transmission resistance, 94
Rolling resistance, 35, 49, 52, 93
Tullberg model, 84-86
coefficient of, 57, 36

S V
Vehicle performance, 92-108
Selection of machinery, 83-88; see also
Machinery selection
Series and parallel arrangements, 82, 83; W
see also Failure probability Weibull distribution, 80, 81
Shear strength of soils, 44, 45 Wheel numeric, 53, 54, 56
Shear stress, 48-51 Winch system, 144

You might also like