You are on page 1of 15

Draught and vertical forces obtained from dynamic

soil cutting by plane tillage tools


A.P. Onwualu
a,*
, K.C. Watts
b
a
Agricultural Engineering Department, University of Nigeria, Nsukka, Nigeria
b
Agricultural Engineering Department, Technical University of Nova Scotia, P.O. Box 1000, Halifax NS B3J 2X4, Canada
Accepted 29 January 1998
Abstract
An understanding of the relationship between tool forces and speed is important in evolving management strategies for optimum
performance. The effect of speed on tillage tool forces were studied experimentally for wide (width=25.4 cm, depth=15 cm) and
narrow (width=5.1 cm, depth=22.9 cm) plane tillage blades operating in a Dystric Fluvisol (silty sand texture) in a soil bin. The
tools were tested at two depths (10 cmand 15 cmfor wide blade, 11.4 cmand 22.9 cmfor narrowblade), two rake angles (458 and
908) and eight speed levels (0.25, 0.5, 0.75, 1.00, 1.25, 1.50, 1.75 and 2.00 m/s). The variables were combined in a 228 factorial
experiment with three replications. The performance of three theoretical models based on the trial wedge approach in predicting
the experimental results was evaluated. The rst model (Model 1), based on Soehne's approach (with modication for the three-
dimensional analysis) assumes that the soil fails in a series of shear planes, forming a wedge that is trapezoidal in shape. The
equilibrium of the wedge boundary forces produce the force required for failure. The second model (Model 2), based on Mckyes'
approach assumes that soil failure is by the formation of a centre wedge anked by two side crescents. Equilibriumof the boundary
forces on the wedge and crescents produce the forces as a function of an unknown failure angle which is obtained by minimizing
the weight component of the total force. Model 3, based on Perumpral's approach assumes the same failure wedge as Model 2
but the total cutting force is minimized instead. Experimental results show that the tool force (draught and vertical force) is a
function of the speed and the square of speed whereas the three models assume it to be a function of the square of speed only.
The models were not very accurate in predicting the experimental results. The average percent deviation of the predicted forces
fromthe observed values were 43%, 40%and 66%for Models 1, 2 and 3, respectively. Thus, Model 2 had more general agreement
with experimental observations. The models were better in predicting the forces (draught and vertical force) for the narrow tool
with average percent deviations of 33%, 28% and 46% for Models 1, 2 and 3, respectively, as compared to 53%, 51% and 85%
for the wide blade. #1998 Elsevier Science B.V. All rights reserved.
Keywords: Tillage; Soil-tool interaction; Soil bin; Draught; Vertical force
1. Introduction
One of the main aims of a good farm manager is to
prepare the soil for planting in the shortest possible
time. This can be accomplished by maximizing the
eld capacity of the tillage implement. The eld
capacity, which is the rate of eld coverage, is the
product of the width and speed of operation. The
choice is, therefore, between operating large equip-
ments at low speed or smaller equipments at higher
Soil & Tillage Research 48 (1998) 239253
*Corresponding author.
0167-1987/98/$ see front matter # 1998 Elsevier Science B.V. All rights reserved.
PI I : S0 1 6 7 - 1 9 8 7 ( 9 8 ) 0 0 1 2 7 - 5
speed. The combination that accomplishes the task in
the shortest time and keeps the power requirements
and accompanying xed and operating costs at a
minimum is usually selected. In making this decision,
the relationship between tool force and speed must be
known.
The relationship between draught and speed has
been reported as linear, second-order polynomial,
parabolic and exponential (Rowe and Barnes, 1961;
Siemens et al., 1965; Stafford, 1979; Swick and
Perumpral, 1988; Gupta et al., 1989; Owen, 1989).
These differences occur as a result of the inertia
required to accelerate soil, effect of shear rate on
shear strength and effect of shear rate on soil-metal
friction, all of which vary with soil type and condition.
For sandy soils, the effect of the inertial forces is more
signicant (Luth and Wismer, 1971). Since inertial
forces increase as the square of the speed, draught
increases as the square of the speed for such soils
(Terpstra, 1977; Owen, 1989). For clay soils, the effect
of shear rate on shear and adhesive strength is more
signicant (Rowe and Barnes, 1961; Wismer and
Luth, 1972). For such soils, draught increases expo-
nentially with speed (Stafford, 1979).
The usefulness of these relationships is in the
development of models which can be used in estimat-
ing draught and hence, power requirement. The sim-
plest of such approaches is the use of regression
equations. Such equations are given by ASAE Stan-
dards D230.4 for a limited range of soil and tool
conditions (Hahn and Rosentreter, 1989). For other
soil and tool conditions not given in this standard, new
regression equations must be developed empirically.
Semi-empirical approaches involve experimental stu-
dies on the effect of shear rate on soil strength.
Relationships developed are then used in the static
theoretical models to scale soil strength instead of
using constant strength parameters (Rowe and Barnes,
1961). Most theoretical models for predicting
dynamic soil-tool interaction are based on the addition
of a velocity component to the static wedge approach.
For both two-dimensional (2-D) and three-dimen-
sional (3-D) analyses, this involves equations that
describe the acceleration of the wedge from zero to
a velocity that enables it to slide up the interface. This
is based on an analysis by Soehne (Gill and Vandern
Berg, 1968) and has been used in different forms for
the 2-D analysis (Rowe and Barnes, 1961; Mckyes,
1985) and the 3-D analysis (Owen, 1989; Gupta et al.,
1989).
Different models based on the wedge approach for
soiltool interaction have been evaluated in the past
for narrow blades operating at very slow speeds, the
so-called passive case (Plasse et al., 1985; Grisso and
Perumpral, 1985). A comparison of their performance
in predicting dynamic soiltool interaction for both
wide and narrow blades has not been done in an
integrated manner. Yet such information is required
for modeling of soiltool interaction.
The objectives of this work were: to develop regres-
sion equations relating the draught and vertical force
of plane wide and narrow tools to speed; modify the 2-
D model of Soehne ( Gill and Vandern Berg, 1968) to
include the 3-D analysis; develop computer pro-
grammes for analysis of soiltool interaction based
on the above model and that by Mckyes (Mckyes and
Desir, 1984) and Swick and Perumpral (1988); and
compare the performance of the models in predicting
experimental observations with respect to the effect of
speed on tool draught and vertical force. Since the
main interest of the study was on the effects of speed,
and in order to keep the experiments within manage-
able limits, it was limited to one soil type (Dystric
Fluvisol) and one soil moisture content (140 g kg
1
).
The moisture content was chosen to coincide with the
normal moisture level for tillage operations for this
soil type.
2. Materials and methods
2.1. Soil bin equipment
The experiments were conducted in the soil bin
facility of the Department of Agricultural Engineer-
ing, Technical University of Nova Scotia, Halifax,
Canada. The facility consisted of a stationary bin,
common carriage that supports either the tool and
penetrometer carriage or the soil processing carriage,
an integrated hydraulic power system, instrumenta-
tion, computer based data acquisition and control
system (Onwualu and Watts, 1989a, b).
The bin was a stationary soil box 7.32 m long,
1.23 m wide and 0.6 m deep, with two rails on top,
one on either side, on which the common carriage was
made to ride. Either the tool carriage or the soil
240 A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253
processing carriage could be placed on top of the
common carriage for any of the operations. The
common carriage and hence any of the carriages
placed on it was pulled in either direction by two
chains driven by the hydraulic power system.
The tool/penetrometer carriage consisted of a frame
with a mounting plate on which a dynamometer,
penetrometer and the tillage tool were mounted.
The soil processing carriage consisted of a frame,
sprayer, rotary tiller, levelling blade and roller. The
motion of all these were powered by the hydraulic
system and controlled through the computer.
The instrumentation consisted of: three rotary
optical encoders attached to the drive shafts for
measurement of the X, Y, Z displacements and
velocities of the tool and penetrometer, an extended
octagonal ring dynamometer for measurement of
draught, vertical force and moment and a penetrom-
eter load cell. The data acquisition and control system
consisted of a personal computer which controlled
ve input/output (I/O) modules through a high speed
digital interface board. The I/O modules included
two encoder interface cards equipped with 20 bit
digital counters, a 12 bit successive approximation
register analog to digital converter (ADC), a 12 bit
digital to analog converter (DAC), and a digital I/O.
The computer acquired data from the penetrometer
load cell and the extended octagonal ring dynam-
ometer through the ADC and that from the encoders
through the encoder interface cards and drove the
actuators through the DAC and the digital I/Os. The
measurement and control system was driven by a
software developed in Quick Basic which ensured
simultaneous acquisition of data and motion control
at a sampling rate of 2.3 kHz.
2.2. Soil description and soil preparation
The soil used for the study was taken from the
Stewiacke area of Nova Scotia, Canada. It has been
classied as Dystric Fluvisol under the FAO/UNESCO
classication system (FAO/UNESCO, 1990). Under
the Canadian system, it is classied as Stewiacke Soil
Series (Webb et al., 1989). Some physical properties
of the soil are shown in Table 1. For each test, a 0.5 m
depth of soil was maintained in the bin. For all tests,
the soil was prepared to an average dry density of
1.5 g/cm
3
, cone index of 0.31 MPa and moisture
content of 140 g kg
1
. These values were averaged
over 25 cm of soil depth. The choice of these values
was informed by the actual average values of soil
properties in the eld during normal eld tillage
operations in the area. The soil was prepared to the
desired density and moisture status by using the
soil processing carriage. To accomplish this, the tiller
was used to pulverize the soil while spraying water
as desired. Then the soil was levelled with the level-
ling blade. Following this, the roller was used to
compact the soil to the desired density in layers. A
special procedure was developed to ensure uniformity
of soil along and across the bin after each soil
preparation.
At the end of each soil preparation, a hydraulically
powered, computer controlled penetrometer mounted
on the tool/penetrometer carriage was used to check
for uniformity at three designated locations on the bin.
The cone index readings were taken every 0.5 cm up
to 30 cm. In addition, soil bulk density and soil
moisture content were measured at these locations
(at 10, 20 and 30 cm soil depth). The locations were
2 m apart along the bin and were selected to check the
soil condition near the entrance of the bin, at the
middle and towards the far end. At each of these
locations, two samples were taken across the bin
(60 cm apart). The locations were chosen so as not
to interfere with actual tillage tests. In checking for
soil uniformity, the soil preparation was repeated if the
Table 1
Soil description
Soil type Stewiacke soil
(Agriculture Canada)
Dystric Fluvisol
(FAO/UNESCO)
Soil texture Silty sand
Clay content (<0.002 mm) 2.5%
Silt content (0.0020.05 mm) 5.0%
Sand content (>0.05 mm) 92.5%
Cohesion (C) 2 kPa
Angle of internal friction 308
Adhesion (C
a
) 7.66 kPa
Angle of soil-metal friction 15.228
Optimum moisture content (dry basis) 122 g kg
1
Maximum dry density 1.73 g/cm
3
Dry density for tests 1.5 g/cm
3
Moisture content for tests 140 g kg
1
A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253 241
soil properties were signicantly different from each
other. More details of these are given elsewhere
(Onwualu and Watts, 1993).
2.3. Treatments and experimental procedure
Tillage tools can generally be classied as wide and
narrow. When the depth of the tool is large compared
to the width, soil movement is in two directions: ahead
and to the sides, and the soil failure problem becomes
three dimensional. Tools in this category (narrow)
include chisels, subsoilers, tines and cultivators. For
other tools, the depth is small compared to the width
and they are called wide blades. Soil movement is only
ahead and so is a 2-D problem. Tools in this category
(wide) include mouldboard ploughs, disc ploughs, and
disc harrows. Based on the works of Payne and Tanner,
1959, a wide blade is dened in this study as that with
aspect ratio (AR) (depth/width) less than 1 and a
narrow tool is dened as that with AR equal to or
greater than 1. The size (depth and width) of the plane
blades used in this study were chosen based on these
principles in order to cover the different classes of
tools.
Tests were conducted using a wide blade (width=
25.4 cm, depth=15 cm) and a narrow blade (width=
5.1 cm, depth=22.9 cm). For the wide blade, the
treatments were two depths (10 and 15 cm), two rake
angles (45 and 908) and eight speeds (0.25, 0.50, 0.75,
1.00, 1.25, 1.50, 1.75, 2.00 m/s). For the narrow blade,
the treatments were two depths (11.4, 22.9 cm), two
rake angles (458 and 908) and eight speeds (0.25, 0.50,
0.75, 1.00, 1.25, 1.50, 1.75, 2.00 m/s). The speeds
were chosen to include actual eld speeds for tillage.
The variables were combined in a 228 factorial
experiment with three blocks or replications. For each
experiment, the desired rake angle, depth and speed
were selected and maintained by the control system.
As the tool carriage and hence the tool moved through
the soil, the measurement system acquired displace-
ment and speed data (X, Y, Z) from the three optical
shaft encoders mounted on the drive shafts. In addi-
tion, force data (draught, F
x
, vertical force, F
y
and
moment, M
y
) were simultaneously and continuously
acquired from an extended octagonal ring dynam-
ometer on which the tool was mounted. Details of
the measurement and control system were given by
Onwualu and Watts, (1989b) and other aspects of the
procedure were given by Onwualu and Watts (1989b)
and Onwualu (1991)
2.4. Simulation models
Different models have been developed for analysis
of soiltool interaction. These include those based on
the trial wedge approach, the stress characteristics
approach, the critical state soil mechanics approach
and the nite element method. Of the four approaches,
the trial wedge approach is the simplest (mathemati-
cally) and so was chosen for this study. Three models
based on the trial wedge approach were selected for
evaluation in this study based on their ability to handle
the effect of speed, simplicity, and ease of program-
ming. The other models based on the same approach
(Osman, 1964; Hettiaratchi et al., 1966; Godwin and
Spoor, 1977) assume curved failure boundary and are
for static analysis (slow moving tools). In addition,
part of the solution depends on the use of charts and
separate experiment for the rupture distance. The three
models were: Model 1 based on Soehne's approach
with some modications (Gill and Vandern Berg,
1968); Model 2 based on McKyes' approach
(Mckyes and Desir, 1984) and Model 3 based on
Perumpral's approach (Swick and Perumpral, 1988).
The soil failure pattern for Model 1 is shown in
Fig. 1(a). The soil is assumed to fail in a series of shear
failure planes which can be approximated to a trape-
zium as shown in Fig. 1(b). For the two-dimensional
case (Fig. 2(a)) as presented by Soehne ( Gill and
Vandern Berg, 1968) and modied by Rowe and
Barnes (1961), the forces acting on the soiltool
system include the weight of the soil wedge (W),
cohesive force on the failure surface due to shearing
(CF
1
), normal component of soil reaction on the fail-
ure surface (N
1
), tangential component of soil reaction
on the failure surface (j
1
N
1
), adhesive force on the
interface (AF
0
), normal component of soil reaction on
the soiltool interface (N
0
), tangential component of
soil reaction on the tool (j
0
N
0
), acceleration force
(F
a
), vertical force (V) and draught (H). The coef-
cient of soilmetal friction (j
0
) is equal to tan c where
c is angle of soilmetal friction while the coefcient of
soilsoil friction (j
1
) is equal to tan c where c is angle
of internal friction of the soil.
Modications were made to the model to enable
it handle 3-D analysis. The concept of the formation
242 A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253
of the centre wedge anked by two side crescents
was used (Mckyes, 1985). The crescents were
assumed to be bounded by straight lines instead of
curves. As the tool moved, a passive condition was
created in the side crescents and so Rankine passive
earth pressure (Terzaghi and Peck, 1967) was
applied to replace the crescents. This resulted in
three forces, namely the normal component of soil
reaction on the side wedge (N
2
), soilsoil frictional
resistance (SF) and resistance to cohesion (CF
2
) as
shown in Fig. 2(b). Each of these forces acts on
both sides of the centre wedge. The force N
2
was estimated according to Rankine's earth pressure
theory as
N
2
=
K
p
dA
2
3
(1)
K
p
=
1 sin c
1 sin c
(2)
SF = j
2
N
2
= N
2
tan c (3)
CF
2
= CA
2
(4)
where K
p
is the coefficient of passive earth pressure,
the bulk density, d the depth of operation, A
2
the
area of the side crescent and j
2
the coefficient of
soilsoil friction or tangent of the angle of internal
friction.
With these additions, by summing forces vertically
and horizontally and equating to zero and eliminating
the normal components of the reactions (N
0
, N
1
), the
draught force was shown (Onwualu, 1991) to be
H =
W
Z

CA
1
2CA
2
2N
2
tan c F
a
Z(sin u tan ccos u)

C
a
A
0
Z(sin c tan c cos c)
(5)
The vertical force (V) becomes
V =
H(cos c tan c sin c)
sin c tan c cos c
(6)
where
Z =
cos c tan c sin c
sin c tan c cos c

cos u tan csin u
sin u tan ccos u
(7)
Fig. 1. Two-dimensional soil failure in front of a tillage tool: (a)
wedge (b) geometrical relationships (Rowe and Barnes, 1961; Gill
and Vandern Berg, 1968) See Table 6 for symbols.
Fig. 2. Free body diagram of the soil tool system for Soehne's
approach (a) 2-D analysis and (b) 3-D analysis. See Table 6 for
symbols.
A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253 243
where: c is rake angle of tool, u is failure angle as
shown in Fig. 1(b), C is cohesion, C
a
is adhesion, Wis
weight of soil wedge, F
a
is acceleration force and A
0
,
A
1
, A
2
are cross sectional areas dened below.
The acceleration force (F
a
) is equal to the resistance
required to bring the block of soil initially at rest to a
speed that ensures that it travels over the tool. This
force is a function of the soil bulk density (), tool
width (b), tool depth (d), tool velocity (v), rake angle
(c) and failure angle (u):
F
a
= bdi
2
sin c
sin(c u)
(8)
The weight of the soil wedge is
W = V
0
(9)
where V
0
= bA
2
is the volume of the wedge.The
failure angle is obtained from passive earth pressure
theory as
u = 45
c
2
(10)
From Fig. 1(b), the various areas are obtained as
A
0
=
bd
sin c
(11)
A
1
=
bd
sin u
(12)
A
2
= d
+
L
0

L
1
L
2
2

(13)
d
+
=
d sin(c u)
sin u
(14)
L
o
=
d
sin c
(15)
L
1
=
d cos(c u)
sin u
(16)
L
2
= d
+
tan c (17)
This model as presented is for 3-D analysis for narrow
tool. To use it for wide blade (2-D), N
2
, SF and CF
2
are
set to zero.
The failure wedge for Model 2 which is credited to
Mckyes is shown in Fig. 3 to consist of a plane centre
wedge, anked by two circular crescents (Mckyes and
Ali, 1977; Mckyes, 1985). The forces involved include
the weight of the soil wedge (), surcharge (q), soil
reaction (R
1
), cohesion (C), adhesion (C
a
) and cutting
force (P). The cutting force (P) per unit width of blade
is given in the form of the universal earth moving
equation, UEE (Reece, 1964) with a dynamic term to
account for speed:
P = gd
2
N

CdN
c
QdN
q
C
a
dN
ca
i
2
dN
a
(18)
where the N-factors are given as:
N

=
r,2d 1 (2r,3b)sin , [ [
cos(c c) sin(c c)cot(u c)
(19)
N
c
=
1 cot u cot(u c) 1 (r,b)sin , [ [
cos(c c) sin(c c) cot(u c)
(20)
N
q
=
(r,d) 1 (r,b)sin , [ [
cos(c c) sin(c c) cot(u c)
(21)
N
ca
=
1 cot ccot(u c)
cos(c c) sin(c c) cot(u c)
(22)
N
a
=
[tan u cot(u c)[ 1 (r,b) sin , [ [
[cos(c c) sin(c c)cot(u c)[[tan u cot c[
(23)
The rupture distance r is given as
r = d(cot c cot u) (24)
The cutting force is obtained by solving the
Fig. 3. Three-dimensional soil failure in front of a tillage tool: (a)
failure wedge (b) force analysis (Mckyes, 1985). See Table 6 for
symbols.
244 A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253
equation
dN

du
= 0 (25)
to obtain the required failure angle u to minimize the
N-factor for weight N

. This failure angle is then used


to obtain the other N-factors and hence P. The draught
(H) and the vertical force (V) are obtained by combin-
ing P with force of adhesion:
H = [Psin(a c) C
a
d cot c[b (26)
V = [Pcos(c c) C
a
d[b (27)
In using the model for 2-D analysis, the angle sub-
tended by the side failure crescent (,) is set to zero.
Model 3 which is based on the approach presented
by Swick and Perumpral (1988) assumes the same
failure wedge and force system as Model 2. An equili-
briumanalysis of the forces actingonthewedge (Fig. 3)
produced the expressions for the cutting force for the
centre wedge (P
1
) and that for the side crescent (P
2
):
where
F
cal
=
C
a
bd
sin c
(30)
W
1
=
bdr
2
(31)
W
2
=
1
6
dr
2
(32)
F
q1
= qbr (33)
F
q2
=
1
2
qr
2
(34)
F
c1
=
cbd
sin u
(35)
F
c2
=
Cdr
2 sin u
(36)
F
a1
= bdi
2
sin c
sin(c u)
(37)
F
a2
=

2
dri
2
sin c
sin(c u)
(38)
where F
ca
is the force of adhesion on the centre
wedge, W the weight of the soil wedge, F
q
the
surcharge force, F
c
the cohesive force, F
a
the
acceleration force. The subscripts 1 and 2 refer to
the centre wedge and the side crescents, respectively.
The angle subtended by the side failure wedge is
obtained from geometrical considerations as
, = sin
1
s
r

(39)
Instead of obtaining s from geometrical considera-
tions, Swick and Perumpral (1988) recommended an
empirical equation which was found to be adequate
after experimental verication:
S = 6.03 0.46r 0.90c (40)
To obtain the failure angle, u the total force P is
minimised with respect to u:
dp
du
= 0 (41)
where
P = P
1
2P
2
(42)
Once the failure angle is obtained, the cutting force is
obtained and hence the draught (H) and vertical force
(V) as before. The model as presented is a 3-D one for
narrow tools. In order to use it as 2-D model, the angle
subtended by the side failure crescent is set to zero.
2.5. Computer implementation
An interactive computer programme was developed
in Microsoft Fortran 77 for simulation of dynamic
soiltool interaction using the three models presented
above. It allows the user, through a menu to analyze 2-
D or 3-D soil failure using any or all the three models.
Results can be obtained in terms of draught and
vertical force as functions of speed for any chosen
tool or soil parameters. The programme was used to
P
1
=
F
cal
cos(c c u) (W
1
F
q1
sin(c u) (F
c1
F
a1
)cos c
sin(c c u c)
(28)
P
2
=
(W
2
F
q2
)sin(c u) sin , F
c2
cos csin , F
a2
cos c[(,,2) sin(,,2)[
sin(c c u c)
(29)
A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253 245
predict the experimental results. Details of the pro-
gramme are given by Onwualu (1991).
2.6. Data analysis
Draught and vertical force data obtained from the
experiments were subjected to analysis of variance
(ANOVA) to determine treatment effects. The
ANOVA was done according to procedures outlined
by Steel and Torrie (1960). Following this, regression
analysis was done to develop the relationships
between the force and speed. Regression models
considered include linear, parabolic, second-order
polynomial and exponential.
The performance of the models in predicting the
experimental results was evaluated by plotting graphs
of the forces (draught and vertical force) vs. speed for
different values of depth and rake angle for the two
blades. In addition, the deviation (%) of the theoretical
results from the experimental ones was obtained as
Deviation(%)
=
Experimental force Theoretical force
Experimental force

100
(43)
For each depth and rake angle combination, the aver-
age deviation was obtained over the eight speed ranges
used in the study. These averages were calculated
using absolute values of the individual deviations.
3. Results and discussion
3.1. Experimental results
Draught and vertical force were signicantly
affected by depth, rake angle and speed at P
0.05
as
shown in Table 2. The interactions were not generally
signicant, showing that the effect of speed did not
depend on depth or rake angle. The main interest in
Table 2
Analysis of variance for tool force
Source of variation df Draught Vertical force
MS F ratio MS F ratio
Wide blade
Block 2 1 395 713 254
*
58 689 41
*
Depth 1 3 959 719 721
*
97 857 69
*
Angle 1 30 554 138 1000
*
32 742 544 1000
*
Speed 7 673 704 123
*
16 836 12
*
Depthangle 1 191 441 35
*
639 450 451
*
Depthspeed 7 1680 0.3 n.s. 1675 1 n.s.
Angle speed 7 55 946 10
*
124 414 88
*
Error 51 5488 1419
Total 95
Narrow blade
Block 2 290 117 104
*
20 000 42
*
Depth 1 12 611 375 1000
*
356 510 75
*
Angle 1 6 017 513 1000
*
7 392 600 1000
*
Speed 7 174 373 63
*
33 590 7
*
Depthangle 1 479 544 171
*
1 008 600 213
*
Depthspeed 7 14 028 5 n.s. 2348 0.5 n.s.
Anglespeed 7 7047 3 n.s. 12 722 3 n.s.
Error 51 2791 4731
Total 95
MS = Mean square.
df = degree of freedom.
*
Significant at P_0.05.
n.s. = Not significant.
246 A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253
this study is speed and so subsequent discussions are
on speedforce relationships.
The forces generally increased with the speed of
operation. Of the possible regression models evalu-
ated, the best model that describes the relationship is a
polynomial of the second degree of the form:
Y = u
0
u
1
i u
2
i
2
(44)
where Y is the draught or vertical force (N), v the tool
speed (in m/s), u
0
, u1 and u
2
the regression coef-
cients given in Tables 3 and 4. The Tables show that
the R
2
values for the model were very high (>0.9) for
both the wide and narrow blade. The relative con-
tribution of v
2
to the model was small as shown by the
coefcients of v
2
compared to v. The results obtained
in this study agree with some earlier reports (Siemens
et al., 1965; Luth and Wismer, 1971; Stafford, 1979;
Owen, 1989) but are at variance with the linear
relationship reported by others (Payne, 1956; Rowe
and Barnes, 1961). The polynomial relationship is
attributed to the combined effect of inertial forces
(square of speed) as the soil slides over the tool and the
effect of rate of shear on tool forces (linear).
3.2. Model prediction: draught
The effect of speed on draught for the wide blade is
shown in Fig. 4 for the two rake angles and two depths
studied. The error bars in the gures refer to 95%
condence intervals for the mean of three replications.
The average percent deviation of the predicted values
from experimental observations are shown in Table 5.
For the wide blade, the average deviation was 62%,
40% and 64% for Models 1, 2 and 3, respectively. The
corresponding values for the narrow blade were 46%,
21% and 37%. For the 458 rake angle, the gure shows
that the models predicted the correct response for
draughtspeed relationship but the values predicted
were generally lower than the experimental results for
the 458 rake angle. This under-prediction at low rake
Table 3
Regression coefficients for the model H=u
0
u
1
vu
2
v
2
for draught force
Depth (cm) Angle (deg.) u
0
P
0.05
u
1
P
0.05
u
2
P
0.05
R
2
Wide blade
10 45 416 0.000 324 0.000 30 0.054 0.99
15 45 721 0.000 280 0.004 9 0.043 0.99
10 90 1029 0.000 878 0.000 153 0.013 0.99
15 90 1683 0.000 630 0.000 78 0.009 0.99
Narrow blade
11.4 45 116 0.000 68 0.023 11 0.079 0.99
22.9 45 553 0.000 195 0.000 13 0.165 0.99
11.4 90 346 0.000 212 0.002 10 0.054 0.99
22.9 90 1110 0.000 308 0.002 13 0.097 0.99
Table 4
Regression coefficients for the model V=u
0
u
1
vu
2
v
2
for vertical the force
Depth (cm) Angle (deg) u
0
P
0.05
u
1
P
0.05
u
2
P
0.05
R
2
Wide blade
10 45 111 0.000 138 0.000 21 0.041 0.99
15 45 141 0.000 261 0.000 65 0.003 0.99
10 90 454 0.000 344 0.001 43 0.105 0.99
15 90 701 0.000 372 0.000 75 0.010 0.99
Narrow blade
11.4 45 38 0.005 91 0.002 21 0.032 0.97
22.9 45 192 0.000 117 0.001 17 0.089 0.99
11.4 90 186 0.000 28 0.055 15 0.028 0.99
22.9 90 378 0.000 166 0.012 49 0.048 0.99
A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253 247
angles may be attributed to at least three reasons,
namely, the assumption that the failure line below the
soil wedge is a straight line instead of a curve, the
bulldozing effect of plane tools at low rake angles,
which is not accounted for by the models and the
assumption that the cutting resistance of the soil is
negligible as assumed by all the models.
For the 908 rake angle, the model based on Soehne's
approach (Model 1) over-predicted the draught while
those based on McKyes' approach (Model 2) and
Perumpral's approach (Model 3) under-predicted the
draught force. The over-prediction by Model 1 actu-
ally starts from about 708 rake angle and that was
probably why Siemens et al. (1965), in using it for
passive analysis recommended that it should not be
used beyond 708. The over-prediction by Model 1 is
attributed to the geometry of the soil wedge assumed
(trapezium) which results in the expressions for the
tool forces having tangent of the rake angle. Since the
tangent of any angle gets abnormally large near 908,
the forces become too large. The present results
reafrm that this model should be used for rake angles
less than 708.
It is noted that although Model 2 assumes the same
failure wedge as Model 3, the results show that the
prediction by Model 3 was lower in magnitude. This
is attributed to the fact that the solution techniques
are different. In Model 2, the failure angle is obtained
by minimising the N-factor for weight while in
Model 3, the failure angle is obtained by minimising
the total cutting force instead of the component for
weight.
The models performed better in predicting draught
for the narrow blade, compared to the wide blade as
shown in Fig. 5. Table 5 shows that whereas the
overall average deviations were 33%, 28% and 46%
for Models 1, 2 and 3, respectively, for the narrow
blade, the corresponding values were 53%, 51% and
85% for the model wide blade. For the 458 rake angle,
Model 2 was generally good in predicting the draught
speed relationship and the other two models predicted
values a little too low. For the 908 rake angle, as was
the case with the wide blade, over prediction of the
draught by Model 1 was obtained. The other two
models had moderate agreement with the experimen-
tal results.
Fig. 4. Effect of speed on draught, wide blade, width=25.4 cm. Soil type= Dystric Fluvisol (silty sand texture), bulk density=1.5 g/cm
3
,
moisture content=140 g kg
1
.
248 A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253
The proportional increase of draught with tool
speed was higher for the narrow tool as predicted
by Models 2 and 3 (Fig. 5). This is obvious from the
relative curvature of the lines in Fig. 5 compared to
Fig. 4. This is attributed to the additional acceleration
force from the side failure crescent in these models.
Since Model 1 does not account for the side failure
crescent even for 3-D analysis, the proportional
increase in draught is similar for wide and narrow
blades.
The three models assume that draught is a function
of the square of speed. The experimental results, as
discussed earlier, indicated that, while this can be
acceptable, a more correct assumption would be that
it is a function of both the speed and the square of the
speed. The conclusion from this is that the accelera-
tion forces in the models should be modied to include
not only the square of the speed but also the speed.
3.3. Model prediction: vertical force
Vertical force prediction by the models as compared
to experimental data for the wide blade is shown in
Fig. 6. The percent deviations are shown in Table 5.
For the wide blade, the average deviations were 44%,
62% and 106% for Models 1, 2 and 3, respectively.
The corresponding values for the narrow blade were
21%, 34% and 56%. For the 458 rake angle and wide
blade, Model 1 predicted values that have moderate
agreement with experimental results although the
model values were generally higher. As was the case
with draught, Models 2 and 3 under-predicted the
vertical force. They were so low for the 458 rake
angle (wide blade) that they were negative instead
of positive. This could be attributed to the nature of the
soil (sandy). Thus for very low rake angles, the
contribution from the weight of the soil wedge was
Table 5
Average percent deviation of predicted forces from experimental observation
a
Blade type Force type Depth (cm) Rate angle (deg) Average % deviation from experiment
b
Model 1 Model 2 Model 3
Wide Draught 10 45 38 48 62
15 45 17 30 56
0 90 57 34 77
15 90 137 48 62
Average for draught force 62 40 64
Wide Vertical force 10 45 47 87 160
15 45 56 93 159
10 90 47 34 56
15 90 16 42 49
Average for vertical force 44 62 106
Average for wide blade 53 51 85
Narrow Draught 11.4 45 25 6 39
22.9 45 25 8 38
11.4 90 26 25 50
22.9 90 106 46 19
Average for draught force 46 21 37
Narrow Vertical force 45 27 46 99
45 13 19 58
90 40 24 44
90 4 47 24
Average for vertical force 21 34 56
Average for narrow blade 33 28 46
Overall deviation 43 40 66
a
Each value is the average of eight speed levels used in the study.
b
Figures are absolute values of the deviations.
A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253 249
smaller than that from adhesion. For the same rake
angle and for the narrow tool, the models showed
moderate but consistently lower agreement with
experimental results (Fig. 7). The vertical force was
no longer negative as obtained with the wide blade
because of the added weight due to the side crescent.
For the higher rake angle (908), the vertical force
was negative for both experimental and model pre-
diction results for the wide and narrow tools (Figs. 6
and 7). The values were negative because as rake angle
increases, vertical force decreases, until at about 608
when it is zero. Beyond this, the tool tends to pull itself
into the soil (suction), thus making the vertical force to
change sign. The models predicted lower values
except for Model 1 which was good in prediction in
some of the cases shown in the gures (wide blade:
d=15 cm, c=908 and most of the predictions for the
narrow tool). Model 2 had moderate agreement with
experimental results for two cases for the narrow tool
(c=458, d=11.4, 22.9 cm) as shown in Fig. 7.
As noted for draught, the prediction of pattern of
variation was not exactly the same for the models and
experimental results. While the experimental results
showed vertical force as a function of the speed and
the square of the speed, the models show it as a
function of the square of the speed only. As noted
earlier, the acceleration equations in the models need
to be modied to account for this.
Table 6. Definition of symbols used in the Figs. 13.
Figure No. Symbol Definition
1 d Tool depth (m)
u Soil failure angle (deg)
d
*
Apparent height of the trapezoidal soil wedge (m)
L
0
Part of length of trapezoidal soil wedge equal to the tool length (m)
L
1
, L
2
Length of leading and failing edge of trapezoidal soil wedge (m)
2 AF
o
Adhesive force on interface (N)
CF
1
Cohesive force on failure surface due to shearing (N)
CF
2
Resistance to cohesion on the side of wedge (N)
F
a
Soil acceleration force (N)
H Draught (N)
SF Soilsoil frictional resistance (N)
V Vertical force (N)
W Weight of soil wedge (N)
j
1
Coefficient of soilsoil friction
N
0
Normal component of soil reaction on the soiltool interface (N)
N
1
Normal component of soil reaction on side wedge (N)
R
0
Soil reaction on the tool surface (N)
N
2
Normal component of soil reaction on the side of the wedge (N)
3 w Tool width (m)
C Cohesion (kPa)
C
a
Adhesion (kPa)
d Tool depth (cm)
P
1
Soil cutting force for centre wedge (N)
P
2
Soil cutting force for side crescent (N)
Soil bulk density (g/cm
3
)
q Surcharge (kPa)
R
1
Soil reaction on the failure plane for centre wedge (N)
R
2
Soil reaction on the failure plane for side crescent (N)
S Width of side failure crescent (m)
r Rupture distance (m)
o Angle of soil metal friction (deg)
c Rake angle (deg)
, Angle subtended by side failure crescent (deg)
c Angle of internal friction of soil (deg)
u Failure angle (deg)
250 A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253
Fig. 5. Effect of speed on draught, narrow blade, width =5.1 cm. Soil type = Dystric Fluvisol (silty sand texture), bulk density=1.5 g/cm
3
,
moisture content=140 g kg
1
.
Fig. 6. Effect of speed on vertical force, wide blade, width=25.4 cm. Soil type = Dystric Fluvisol (silty sand texture), bulk density=1.5 g/
cm
3
, moisture content=140 g kg
1
.
A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253 251
3.4. Overall model prediction
The average deviations of the predicted values from
the experimental ones varied from 6% to 160% as
shown in Table 5. The overall performance of the
models in predicting the experimental results for
draught and vertical force taken together is reected
in the overall deviations shown in Table 5 to be 43%,
40% and 66%, respectively. Thus, Model 2 had more
general agreement with the experimental observa-
tions.
4. Conclusions
Based on the experimental observation, tool force
(draught and vertical force) is a function of speed and
the square of speed. However, the three models eval-
uated in this study assume that the tool force is a
function of the square of the speed only. The models
should be modied to account for this. The three
models under-predicted the tool forces vs. speed
relationship for the wide blade. Moderate agreement
between the predicted and observed values were
obtained for the narrow tool in general for all the
three models. However, Model 1 (based on Soehne's
approach) tends to over-predict draught at 908 rake
angle for both 2-D and 3-D analysis. Model 3 (based
on Perumpral's approach) generally under-predicted
the relationship between tool forces and speed. Model
2 (based on McKyes' approach) had more general
agreement with experimental observations, especially
for the narrow tool when compared with the other
models.
Acknowledgements
The authors are grateful to the technical staff of
Department of Agricultural Engineering, Technical
University of Nova Scotia for equipment fabrication.
These include J. Visers, J. Godwin, C. Wade and A.
Murphy. The wiring of electrical and electronic com-
ponents of the soil bin was by G. Jollimore and G.
Yourick. Funding for the project is by the CIDA-
TUNS-UNN Linkage agreement.
Fig. 7. Effect of speed on vertical force, narrow blade, width=5.1 cm. Soil type = Dystric Fluvisol (silty sand texture), bulk density=1.5 g/
cm
3
, moisture content=140 g kg
1
.
252 A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253
References
FAO/UNESCO, 1990. Soil Map of the World, Revised Legend.
Food and Agricultural Organization of the United Nations
(FAO) and United Nations Educational and Scientific Organi-
zation (UNESCO).
Gill, W.R., Vandern Berg, G.E., 1968. Soil dynamics in tillage and
traction. Agriculture Handbook No. 316. Agricultural Research
Service, United States Department of Agriculture (USDA). pp.
126140.
Godwin, R.J., Spoor, G., 1977. Soil failure with narrow tines. J.
Agric. Eng. Res. 22(4), 213228.
Grisso, R.D., Perumpral, J.V., 1985. Review of models for
predicting the performance of narrow tillage tools. Trans.
Am. Soc. Agric. Eng. (ASAE). 28(4), 10621067.
Gupta, P.D., Gupta, C.P., Pandey, K.P., 1989. An analytical model
for predicting draft forces on convex-type cutting blades. Soil
Till. Res. 14, 131144.
Hahn, R.H., Rosentreter, E.E. (Eds.), 1989. ASAE Data D230.4.
Agricultural machinery management data. In: American
Society of Agricultural Engineers (ASAE) Standards. 36th
ed., ASAE, St. Joseph MI 49085. 61 pp..
Hettiaratchi, D.R.P., Witney, B.D., Reece, A.R., 1966. The
calculation of passive pressure in two-dimensional soil failure.
J. Agric. Eng. Res. 11(2), 89107.
Luth, H.J., Wismer, R.D., 1971. Performance of plane cutting
blades in sand. Trans. Am. Soc. Agric. Eng. (ASAE) 14(2),
255262.
Mckyes, E., 1985. Soil Cutting and Tillage. Developments in
Agricultural Engineering, vol. 7, Elsevier, Amsterdam.
Mckyes, E., Ali, O.S., 1977. The cutting of soil by narrow blades. J.
Terramech. 14(2), 4358.
Mckyes, E., Desir, F.L., 1984. Prediction and field measurements
of tillage tool draft and efficiency in cohesive soils. Soil Till.
Res. 4, 459470.
Onwualu, A.P., 1991. Tillage tool factors affecting sandy soil
interaction with plane blades in a soil bin. Ph.D Thesis,
Technical University of Nova Scotia, Halifax, Canada.
Onwualu, A.P., Watts, K.C., 1989a. Development of a soil bin test
facility. American Society of Agricultural Engineers Paper No.
89-1106, St. Joseph, MI 49085-9659.
Onwualu, A.P., Watts, K.C., 1989b. Control system for a
hydraulically powered soil bin. American Society of Agricul-
tural Engineers Paper No. 89-1570, St. Joseph, MI, 49085-
9659.
Onwualu, A.P., Watts, K.C., 1993. Real time measurement of cone
index in a soil bin. Comput. Electron. Agric. 9, 143157.
Osman, M.S., 1964. The mechanics of soil cutting blades. J. Agric.
Eng. Res. 9(4), 313328.
Owen, G.T., 1989. Subsoiling forces and tool speed in compact
soils. Can. Agric. Eng. 31(1), 1520.
Payne, P.C.J., 1956. The relationship between the mechanical
properties of soil and the performance of simple cultivation
implements. J. Agric. Eng. Res. 1(1), 2350.
Payne, P.C.S., Tanner, D.W., 1959. The relationship between rake
angle and the performance of simple cultivation implements. J.
Agric. Eng. Res. 4(4), 312325.
Plasse, R., Raghavan, G.S.V., McKyes, E., 1985. Simulation of
narrow blade performance in different soils. Trans. Am. Soc.
Agric. Eng. 28(4), 10071012.
Reece, A.R., 1964. The fundamental equation of earth moving
mechanics. Proc. Symp. on Earthmoving Machinery, Institution
of Mechanical Engineers, pp. 1622.
Rowe, R.J., Barnes, K.K., 1961. Influence of speed on elements of
draft of a tillage tool. Trans. Am. Soc. Agric. Eng. 4, 5557.
Siemens, J.C., Weber, J.A., Thornburn, T.H., 1965. Mechanics of
soil as influenced by model tillage tools. Trans. Am. Soc.
Agric. Eng. 8(1), 17.
Stafford, J.V., 1979. The performance of a rigid tine in relation to
soil properties and speed. J. Agric. Eng. Res. 24, 4157.
Steel, R.G.D., Torrie, J.H., 1960. Principles and Procedures of
Statistics With Special Reference to the Biological Sciences.
McGraw-Hill, NY.
Swick, W.C., Perumpral, J.V., 1988. A model for predicting soil-
tool interaction. J. Terramechanica 25(1), 4356.
Terpstra, R., 1977. Draught forces of tines in beds of glass spheres.
J. Agric. Eng. Res. 22, 135143.
Terzaghi, K., Peck, R.B., 1967. Soil Mechanics in Engineering
Practice. Wiley, NY, pp. 192206.
Webb, K.T., Duff, J.P., Langille, D.R., 1989. Soils of the Cobequid
Shore Area of Nova Scotia. Report No. 23, Nova Scotia Soil
Survey. Agriculture Canada and Nova Scotia, Department of
Agriculture and Marketing.
Wismer, R.D., Luth, H.J., 1972. Performance of plane soil
cutting blades in a clay. Trans. Am. Soc. Agric. Eng. 15(2),
211216.
A.P. Onwualu, K.C. Watts / Soil & Tillage Research 48 (1998) 239253 253

You might also like