You are on page 1of 19

2231

A simplified dynamic model for front-end loader design


M D Worley‡ and V La Saponara∗
Department of Mechanical and Aeronautical Engineering, University of California, Davis, California, USA

The manuscript was received on 01 April 2007 and was accepted after revision for publication on 09 May 2008.

DOI: 10.1243/09544062JMES688

Abstract: The front-end loader is an indispensable machine for the off-road construction equip-
ment industry. It is a classic example of a working machine with complex interactions between
its subsystems (hydraulic, mechanical, and electrical). Dynamic models of the full-scale vehicle
coupled with event-based operator models are currently used to help quantify the overall system
performance, efficiency, and operability. However, these models are complex and not always nec-
essary to characterize the response of individual subsystems. There is great value added to the
design process – especially in prototyping of new vehicle platforms – in development of simpler
models that can quickly and accurately define first-order measures of system loads and per-
formance. This paper presents a subscale dynamic model, which isolates the boom and bucket
manipulator systems of the front-end loader for the purpose of design load characterization. The
model includes state equations governing the hydraulic dynamics across the control valves and
in the cylinders, as well as soil–tool interaction loads (passive earth loads) at the bucket cutting
edge. The governing equations of motion for the multi-rigid body model of the bucket linkage
are developed using Kane’s method. The proposed model is intended to accelerate the structural
design and analysis of the boom and bucket linkage subsystems and may yield useful information
for optimization purposes. The output from the dynamic simulation is compared with the field
test data of the machine.

Keywords: front-end loader, dynamic model, Kane’s method

1 INTRODUCTION complex and may not always be necessary for char-


acterizing the response of individual subsystems for
The front-end loader is a classic example of a working the purposes of design. The development of accu-
machine that has complex, often non-linear, interac- rate, simplified models that accelerate the design
tions among its subsystems (hydraulic, mechanical, process – especially in prototyping of new vehicle plat-
and electrical) [1]. Full-scale vehicle dynamic mod- forms – are of great value to the analyst, and provide
els and event-based operator models represent the the motivation for this study.
current state-of-the-art dynamic simulations, carried This paper proposes a new subscale dynamic model
out with the purpose of quantifying the overall per- for the boom and the bucket manipulator of the front-
formance, efficiency, and operability of the system. end loader, with the goal of adequately characterizing
However, the correct simulation of these factors is the subsystem response for design and optimization
complicated by the variability of the machine’s work purposes. The interaction of the subsystem with the
environment and task profile. While the full-scale sim- rest of the machine via the hydraulic control sys-
ulation may offer the best insight in the machine’s tem is modelled using state equations governing flow
overall performance and efficiency, these models are across the control valves and in the hydraulic cylin-
ders. The interaction of the subsystem with the work
environment via passive earth loads on the digging
∗ Corresponding author: Department of Mechanical and Aeronau- tool (i.e. bucket) is modelled using soil–tool reaction
tical Engineering, University of California, One Shields Avenue, loads that make heavy use of earth moving equa-
Davis, California 95616, USA. email: vlasaponara@ucdavis.edu tions developed by Terzaghi [2] and McKeys [3]. The
‡ Now at: DJH Engineering Centre, Inc., Holladay, Utah, USA.
equations of motion for the multi-rigid body model

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2232 M D Worley and V La Saponara

of the manipulator system are obtained using Kane’s The remainder of this paper is organized as follows.
method. Section 2 provides a brief, general introduction to
Some of the primary assumptions of the model the composition, operation, and application of the
should be highlighted. First, it should be noted that front-end loader. Section 2.1 provides a descrip-
the proposed model is limited in application to the tion of the particular boom and bucket mecha-
digging phase of the front-end loader’s work cycle. nism modelled. Section 2.2 outlines the simplified
Though this is certainly not a comprehensive cross- hydraulic system operation and equations govern-
section of the structural loading experienced by the ing the hydro-mechanical dynamics of the actuator
entire machine, it is the primary loading cycle for cylinders. Section 2.3 describes the interaction forces
the boom and bucket manipulator subsystem and is between the bucket and the dirt pile. A review of
therefore of primary interest. Next, it should be noted the literature revealed extensive discussion regard-
that the proposed model ignores the pitch/bounce ing soil–tool interaction models but little information
dynamics of the tires, owing to the fact that during regarding a detailed procedure for applying them to
the digging operation the vehicle speeds are gener- wheel loaders and sloped pile digging. As such, this
ally low. In more general models that wish to capture section concerning bucket–soil interaction is exhaus-
vehicle dynamics during transport or articulation, the tive in detail, hoping to document the procedure
tire–terrain interaction dynamics cannot be ignored. used here. Section 3 provides information concerning
Lastly, as indicated above, the proposed model ignores assumptions about operator input (cylinder actua-
the elastic behaviour of the boom, bucket, and links, tion). In section 4, the simulation results are correlated
opting to model the system using rigid bodies. The and compared with the field data gathered on an actual
goal of the proposed model is not to mirror the machine. Discussion about results and areas of future
measured load–time history from an actual machine, improvement is also included. Section 5 provides a
which inexorably leads one to build more and more brief summary of the paper, drawing conclusions
complicated machine and operator simulation mod- about the model and its applicability.
els. Rather, the goal of this model is to aptly predict
peak loads during the primary work cycle (dig cycle).
The elastic response of the bucket, boom, and link 2 THE FRONT-END LOADER MODEL
structures, due to either symmetric or asymmetric
application of these peak loads, is expected to be Figure 1 shows a generic machine level description of
captured in subsequent structure and finite element a four-wheel articulated-steer front-end-loader. The
analyses (FEA). articulation axis is the pivot axis between the forward

Fig. 1 General global (planar) schematic used for four-wheel articulated front-end loader model

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2233

and rear frames. With the exception of the front axle by Hemami and Daneshmend [7], Sarata et al. [8],
and its connecting driveshaft, the powertrain system and Scheding et al. [9]. In this paper, the machine
components are located in the rear of the machine. The is assumed fixed in the inertial frame and the
operator cabin is typically mounted on the rear frame, bucket mechanism analysed as a planar mechanism
but in a location (forward and above the articulation grounded in the forward frame. In reality, forces on the
axis) that allows maximum visibility of the working bucket during loading/digging are created, in part, by
tool (e.g. a loading bucket). The forward frame pro- global motion (penetration) of the vehicle into the ter-
vides a platform for the boom and bucket manipulator rain. A method for estimating these loads and applying
linkage, which are collectively labelled in this study them to the bucket as time-varying forces is given later.
as the bucket mechanism. Several bucket manipulator This paper is concerned primarily with the digging
linkage types exist for controlling bucket orientation. phase of the loader’s work-cycle and developing a
Also, it should be noted that the tool at the end of the simplified model.
boom need not necessarily be a loading bucket – the
versatile front-end loader is increasingly being used
and designed to handle a variety of applications. How- 2.1 Bucket mechanism
ever, the loading bucket remains the traditional and
primary work tool for the wheel loader. The bucket mechanism considered for this study is
There are typically three sets of hydraulic cylin- shown in Fig. 2. It has a relatively new bucket manip-
ders and associated systems for kinematic control: ulator linkage type, introduced by John Deere in 2004
(a) steering cylinders (not shown in Fig. 1) for vehi- on its Powerllel™ machines, and was selected because
cle articulation; (b) lift cylinder(s) for controlling the of interest in its unorthodox design: the centre pivot
boom position, and (c) tilt cylinder(s) for controlling joint of the bellcrank is not affixed to the boom
bucket orientation. Fluid pressure and flow in these structure as in other linkage designs. The bellcrank
systems are delivered by hydraulic pumps, which are pivots on a separate link, labelled here as the ‘Y-link’
in turn powered by the vehicle’s engine. In fact, as because the shape of the part’s actual design resem-
all onboard power generation ultimately originates in bles a ‘y’. The Y-link pivots in the forward frame.
the combustion chamber of the engine, this available For clarity, the underlying kinematic geometry that
power must be distributed to, and balanced among defines the boom structure and lift cylinder(s) is shown
the machine’s electrical, hydraulic, and powertrain using dotted lines. Similarly, dashed lines indicate the
demands. Detailed discussion regarding the power geometries of the bucket manipulator linkage and tilt
balance between the hydraulic and powertrain sys- cylinder.
tems can be found in references [1] and [4] to [6]. Figures 3 and 4 show the boom and bucket link-
Furthermore, all system control ultimately originates age systems separated, and define the geometrical
with the operator located in the cabin, who controls (kinematic) values required for the overall mechanism
the subsystems using various levers, pedals, switches, model. The mechanism has ten rigid bodies and two
etc. Dynamic operator models have been developed to degrees of freedom. The equations of motion for the
capture this ‘human element’ in the overall system per- system may be derived with the aid of any commer-
formance of the machine [1, 4, 5]. It is recognized that cially available symbolic manipulator. In this case, the
current state of the art in dynamic simulation of these program Autolev [10] was used to generate and solve
machines employ full-scale vehicle models and event- the equations. This program is especially suited for
based operator models to quantify the overall system employing Kane’s method [11]. Due to their length,
performance, efficiency, and operability. However, for the equations are not included here. However, a brief
the purposes of this study, simplified operator logic description of their set-up is appropriate.
and hydraulic system models suffice to simulate the The trivial set of generalized speeds is chosen as
machine operations considered. This will be discussed shown below
in more detail later in this paper.
N
As suggested by Fig. 1, a comprehensive planar d
U1 = (LLCYL ) = L̇LCYL = vLCYL_ROD (1)
description of the machine dynamics would include dt
translation and pitch motion of the vehicle, also tak- N
d
ing into account the compliance of the tires. More- U2 = (LTCYL ) = L̇TCYL = vTCYL_ROD (2)
dt
over, general loader operations include turning and N
d
transporting the payload, often over uneven terrain. U3 = (φLCYL ) = ωLCYL (3)
Properly constructing a vehicle model to handle a dt
N
wide variety of loading scenarios obviously requires d
U4 = (φTCYL ) = ωTCYL (4)
increasing generality and complexity. Examples of var- dt
N
ious approaches and levels of detail to the modelling d
of loader dynamics can be found in earlier work U5 = (φBOOM ) = ωBOOM (5)
dt

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2234 M D Worley and V La Saponara

Fig. 2 Schematic of bucket mechanism (and attached bucket) used for this study

N
v J11_GDLK = N v J11_BOOM (13)
N
v J12_BKT
= v
N J12_BOOM
(14)

Equations (11) to (14) represent eight planar motion


constraints, dictating continuity of velocities for joint
J5 (belonging to lift cylinder and boom), joint J7
(belonging to bellcrank and Y-link), joint J11 (belong-
ing to guide link and boom), and joint J12 (belonging
to bucket and boom). Substituting the generalized
speeds U1 –U10 in equations (11) to (14) and solving
for U1 and U2 as the independent generalized speeds
(i.e. the system degrees of freedom), one obtains the
necessary dependencies for the generalized speeds
U3 –U10 .
Fig. 3 Kinematic variables of the boom system The contributing active and inertial forces in the
system include: (a) the weights and inertial loads of
N
d the linkage bodies (mass) due to gravity and motional
U6 = (φBLCRK ) = ωBLCRK (6)
dt acceleration; (b) the lift cylinder force acting on the
N
d boom at joint J5; (c) the tilt cylinder force acting on the
U7 = (φBKTLK ) = ωBKTLK (7) bellcrank at joint J8; and (d) the reactive forces on
dt
N the bucket tool due to digging operations. Therefore,
d
U8 = (φBKT ) = ωBKT (8) the planar velocities and accelerations of the system
dt bodies and these idealized points of active loading are
N
d constructed and used to determine the constrained
U9 = (φGDLK ) = ωGDLK (9)
dt partial velocities and partial angular velocities. Dot
N products of these partial velocities with the corre-
d
U10 = (φYLINK ) = ωYLINK (10) sponding active and inertial forces produce expres-
dt
sions for the generalized active and inertial forces,
The following set of motion constraints apply to the which comprise the equations of motion [11].
system The following sections provide the details for the
forces generated by the hydraulics as well as the pas-
N
v J5_LCYL = N v J5_BOOM (11)
sive earth pressure loads generated on the bucket
N
v J7_YLINK
= v
N J7_BLCRK
(12) during digging.

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2235

Fig. 4 Kinematic variables of the bucket linkage system

2.2 Hydraulic dynamics The equations governing pressure inside of the cylin-
ders are given by
Figure 5 shows a schematic of the simplified hydraulic
system model used to actuate the tilt and lift cylin- βF
ders. The hydraulic pump is considered as a positive ṖROD = (QROD + AROD vROD ) (17)
VolROD
displacement device that maintains a supply pressure
PS a fixed P above the cylinder load pressure, unless and
the system relief pressure is reached. Fluid flow to
βF
the cylinders is regulated by turbulent flow across the ṖPIS = (QPIS − APIS vROD ) (18)
control valve orifice. VolPIS
The flow control valves and cylinders are modelled where VolROD and VolPIS are the instantaneous volumes
as in references [12] and [13], with cylinder extension of the rod-side and piston-side chambers, respec-
designated as the positive convention. Therefore, flow tively; βF is the bulk modulus of the hydraulic fluid;
into the piston-side volume of the cylinder and flow AROD and APIS are the fluid areas of the rod-side and
out of the rod-side volume are respectively given by piston-side cross-sections of the cylinder, respectively;
 vROD is the velocity of the rod (positive for extension).
2|PS − PPIS | Equations (15) to (18) are solved at every timestep
QPIS = AVALVE C1 sign(PS − PPIS ) (15) in the dynamic simulation of the bucket mechanism.
ρhyd

and

2|PROD − PE |
QROD = AVALVE C2 sign(PROD − PE ) (16)
ρhyd

where AVALVE is the exposed control valve orifice area,


which is a function of the throw distance of the con-
trol lever by the operator and, moreover, the following
holds:
(a) For cylinder extension: C1 = C1in ; C2 = C2out ; PS =
pump (supply) pressure; PE = tank (exhaust)
pressure;
(b) For cylinder retraction: C1 = C1out ; C2 = C2in ; PS =
tank (exhaust) pressure; PE = pump (supply)
pressure. Fig. 5 Schematic of simplified hydraulics system

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2236 M D Worley and V La Saponara

The operator input and hydraulic pump assumptions (penetration) into the pile as well as the bucket lift and
determine the flowrates for the cylinders while the sys- rotation due to actuation of the lift and tilt cylinders,
tem dynamics determine the velocity of the cylinder respectively. The path is not necessarily known a priori
rods. Time integration of equations (17) and (18) pro- and will depend ultimately on operator commands.
duce the pressures in the cylinders. Finally, the net As pointed out in references [16] and [19], the ‘best’
cylinder forces applied to the system are calculated as or most productive path may vary from cycle to cycle
due to the variability of forces in the digging process
Fcyl = APIS PPIS − AROD PROD (19) created by the soil and/or rock properties.
Hemami [15] describes the bucket–soil interaction
as consisting of five force components:
2.3 Digging forces on the bucket
One of the challenging aspects in modelling the
(a) f1 , the weight of the loaded material;
behaviour of a wheel loader – or any earthmov-
(b) f2 , the resistance force created if the bucket motion
ing equipment – is capturing the forces produced
causes it to compress the dirt pile;
during digging operations. These forces are primar-
(c) f3 , the net friction force between tool and bucket;
ily dependent on the properties of the soil or rock
(d) f4 , the digging resistance acting at the bucket
that is being excavated, which can vary drastically
cutting edge;
depending on size, composition, temperature, mois-
(e) f5 , the inertial force required to accelerate the
ture content, compaction, etc. This variability, how-
accumulated mass.
ever, should and does pose a greater challenge to
efforts for autonomous excavation and productiv-
ity studies, since soil conditions can greatly alter In attempting to describe a path of minimal energy
the method of scooping required to achieve opti- input, he suggests that such a path should cause force
mal performance [14]. For the purposes of struc- f2 to be zero. This point might be debatable, however,
tural design and baseline comparison studies, the as it is at least conceivable that this force could be used
order of magnitude predictions for standard design beneficially in the loading process: if the line of action
cases should suffice. Based on previous work by of force f2 is such that it creates a positive moment
Hemami [15], Singh [16], and Ericsson and Slatten- (with respect to the machine frame M1 , M2 , M3 ) about
gren [17], it appears possible to achieve reasonable the bucket hinge pin (joint J12 in Fig. 3), then this force
estimates of the interaction forces between the cut- is aiding the tilt cylinder in curling the bucket and
ting tool and soil. Furthermore, a two-dimensional payload. This point is made only to ask: ‘Does such a
model is adequate for the geometries of most bucket beneficial scenario exist?’ That is, since such a force f2
tools [16]. Their methods utilize the empirical and ana- would require a proper combination of tractive effort
lytical studies in soil mechanics done by Terzaghi [2], and bucket orientation during the scooping process,
Bekker [18], and McKeys [3], to name a few. Singh [16] it becomes a matter of weighing the additional energy
provides a good synopsis of basic soil mechanics and consumed in the powertrain to the energy saved in the
associated literature. hydraulic cylinder(s).
Figure 6 shows a diagram of the bucket tool dig- The authors believe that all the five force compo-
ging/scooping in a dirt pile. The path of the bucket cut- nents described are included in this proposed model.
ting edge is determined by the global vehicle motion However, they exist in a form more consistent with the
soil wedge model (originally proposed by Coulomb in
1772 and revisited by Terzaghi [2]) detailed by McK-
eys [3]. As previously pointed out, the nature and
activation of these loads may change during the dig-
ging process due to the variability in soil properties,
changes in the contact interface between the bucket
and terrain, and the variability in the operator reac-
tion/preference. To simulate loads imparted to the
bucket tool throughout the digging process requires
making assumptions for all the three. The inaccuracy
due to nominal soil properties is considered here as
unavoidable. Real time, in situ measurements of the
soil properties appear to be the only way to accurately
develop these parameters [16]. The items concern-
ing changing tool–soil contact and operator reaction
are considered interconnected and are thus treated
Fig. 6 Sketch of bucket tool digging in work pile together in the following discussion.

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2237

Fig. 9 Force diagram of bucket tool

orientation α is altered. Inertial effects on the soil


Fig. 7 Three phases of the digging operation mass have been included. At each timestep, a mass
m (shown crosshatched in Fig. 8) is being added to the
bucket and must be accelerated to match the speed of
the soil wedge.
Due to the relative motion between the wedge and
the bucket tool and referring to Fig. 8, the speed V  of
the soil mass m is related to the average speed V of the
bucket tool cutting plane as follows [3]

 
motion of soil mass, m χ 1
V = =
time cos β t

where
Fig. 8 Force diagram of soil wedge
χ = χ  + ε = χ  (1 + tan β cot ρ) (20)

The sketch in Fig. 7 describes the three phases of


Therefore
bucket loading that are assumed in this study. Figures 8
and 9 describe the interaction loads assumed between
the bucket, the soil wedge, and the bulk dirt pile. Not χ V
all these forces are necessarily active at the same time V = =
t cos β(1 + tan β cot ρ) cos β(1 + tan β cot ρ)
during any given phase. Detailed descriptions of active (21)
loads and assumptions for each phase are given in the
following sections.
At this point, a discussion of the soil wedge model where
and associated forces is required. Referring to Figure 8,
force F is the combined net normal and frictional force  
motion of bucket tool χ 1
acting between the soil wedge and tool along bound- V = =
time cos β t
ary AB, acting at the angle δ of soil-to-metal friction.
Force Ca LT is the force due to cohesion between the
soil and metal along boundary AB. Force R is the com- Therefore, the inertial force FA , acting in opposite
bined net normal and frictional force acting between direction to V  , can be written as
the soil wedge and ‘undisturbed’ soil along boundary
BC, acting at the angle of internal shearing resistance V γ · χ · dw
for the soil, φ. Force CLF is the force due to the (appar- FA = m = V = V  γ · V · dw
t t
ent) cohesion of the soil along the failure boundary BC.
γ · V 2 · dw
Load q is the surcharge pressure. The angle ρ = θ − α is = (22)
the rake angle of the cutting edge (commonly referred cos β(1 + tan β cot ρ)
to as the ‘bolt-on-cutting edge’, hence the use of the
abbreviation BOC in Figs 9 and 11) relative to the dirt where γ is the total soil density. Continuing with the
pile surface. This angle changes as the cutting edge method presented by McKeys [3], the unknown R can

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2238 M D Worley and V La Saponara

be eliminated and solve for the force F as kC and kφ are the cohesive and frictional moduli of
 deformation, respectively. It should be observed that
F = γ · g · d 2 · Nγ + C · d · NC + Ca · d · NCa the value for b has considerable influence on the mag-
 nitude of force P and the resulting depth of penetration
+ q · d · N q + γ · V 2 · d · Na · w (23)
prior to initiating wheel slip. This, in turn, influences
the initial payload at the start of the scooping phase
where w is the width of the bucket tool cutting edge
(digging ‘Phase 2’, described in section 2.3.2). Figure 10
and N γ , NC , NCa , Nq , and Na are given by
shows a close-up of a typical cutting edge for a loader
(cot ρ + cot β) · [sin θ cot(β + φ) + cos θ] bucket.
Nγ = (24) The chamfer profile makes straightforward use of
2 · cos(ρ + δ) + sin(ρ + δ) cot(β + φ)
the cutting edge thickness for the dimension b. It is
1 + cot β cot(β + φ) reasoned that using the full thickness is incorrect, as
NC = (25)
cos(ρ + δ) + sin(ρ + δ)cot(β + φ) the soil tends to slide up and off the chamfer slope.
On the other hand, it is also probably not correct to
1 − cot ρ cot(β + φ) use the smaller dimension b1 since the weight of the
NCa = (26)
cos(ρ + δ) + sin(ρ + δ) cot(β + φ) soil above provides constraint to this process. Several
Nq = 2Nγ (27) iterative studies for the penetration phase were run to
help identify a reasonable value for b. The value that
tan β + cot(β + φ) best simulated the penetration depth and pressure
Na = (28) transition from digging ‘Phase 1’ to digging ‘Phase 2’
[cos(ρ + δ) + sin(ρ + δ) cot(β + φ)]
· (1 + tan β cot ρ) (discussed in next section) was determined to be
approximately one-half of the cutting edge thickness.
Noting that LT = d/sin ρ and LF = d/sin β, it is seen Further discussion regarding the setting of b may be
that the loads F and Ca LT are functions of the follow- found in the next sections on the three digging phases.
ing: (a) the soil and soil–tool properties, i.e. δ, φ, γ , C, To determine when and how the aforementioned
and Ca ; (b) the kinematics of the bucket, i.e. ρ, d, and forces are assumed to be active, a more detailed dis-
V ; and (c) the unknown angle β, which is found by cussion regarding the three phases of digging and their
minimization of N γ with respect to β [3]. respective assumptions is appropriate. These three
One disadvantage of the wedge model is that it is phases are Phase 1 ‘crowding’, Phase 2 ‘bucket filling’,
meant to model steady-state loading of soil under and Phase 3 ‘rollback’.
imminent failure. In reality, the compressive load F
builds over a period of time, drops as the soil fails 2.3.1 Phase 1: crowding
along boundary BC, and then builds again until the
next failure. This produces a saw-like profile for the The bucket tool penetrates into the pile at an initial
load F , whereas F is constant in the wedge model. Nev- rake angle ρ0 . The cutting edge motion is due only to
ertheless, the model provides reasonable results with tractive effort creating forward translation of the vehi-
considerable simplicity. cle. In reality, some degree of soil cutting occurs owing
The bucket is assumed to have three points where to the bucket’s sloping profile, and the bucket is being
active loading may occur: the mass centre of the loaded by the developing soil wedge (creating forces
bucket, the mass centre of the developing payload, and F and Ca LT , Fig. 8). However, since the cutting edge is
the cutting edge of the bucket. Referring to Fig. 9, loads supported by the soil underneath it, the components
F and Ca LT are applied to the bucket tool by the soil
wedge. Moreover, a load P and possibly loads Fsupport
and Ca LBOC are applied to the bucket by the dirt pile.
The load P is the penetration load that exists when-
ever the cutting edge is forced further into the dirt
pile. The magnitude of P is determined as the product
of the face area of the bucket cutting edge and the pres-
sure acting on this area. This pressure is approximated
using the empirical load-sinkage formula proposed by
Bekker [18]
 
kC
p= + kφ z n (29)
b

where z is the sinkage depth, n is the exponent of defor-


mation, b is the width of the penetrating surface, and Fig. 10 Close-up of cutting edge

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2239

of F and Fsupport (Fig. 9) that are normal to the cut- manner that allows him/her to maintain this speed
ting edge cancel. Equation (29) is used to calculate the (approximately). This assumption is made based on
penetration force P, where the value z is estimated observations of the loading process and with the
using the depth d. For all phases, the terrain loads intent to provide some autonomy to the operator
acting on the bucket are transformed into the loads commands in Phase 2. The critical transition speed
FV_P3 , FH_P3 , and MP3 , which are, respectively, the ver- assumed in this study is 0.5 m/s, after which a main-
tical force component, horizontal force component, tenance speed range of 0.25–0.5 m/s is set as the
and moment, and act at the cutting edge (i.e. desig- target parameter.
nated as point P3) with respect to the cutting edge 2. If the soil is very loose and non-cohesive, the opera-
frame (Fig. 11). tor may penetrate deeper than desired or necessary
Based on the assumptions aforesaid, the expressions following rule 1. Instead, it is assumed he/she
for each of these forces during Phase 1 are would begin Phase 2 (discussed in the next section)
after reaching some nominal or critical depth.
FV_P3 = F cos δ − Fsupport cos δ = 0 (30) Though somewhat arbitrary, this critical depth is
determined in the following way. A critical or maxi-
FH_P3 = P + 2F sin δ + Ca (LBOC + LT ) · w (31)
mal length LTcrit of tool–soil engagement is assumed
LT to define the point where the bucket profile disrupts
MP3 = F cos δ (32)
2 the flow of soil in the wedge. The maximum length
assumed here is the curve length shown in Fig. 11.
Consideration was given as to how this phase should
end. Most operators do not simply crowd the pile until As part of the simplifications of this study, the state
they lose traction as this represents wasted effort – it equations for the powertrain response have been
generates in excess the force that Hemami [15] labelled omitted. Instead, the forward velocity of the vehicle
as f2 . However, sufficient penetration is required to is governed by the following linearized load–velocity
ensure a full bucket. Therefore, the following simpli- relationship
fying assumptions have been made:
Vgear1
1. It is assumed that the operator desires to keep the VFWD = Vgear1 − H (33)
Rimpullmax
machine moving forward at a relatively constant
speed throughout the majority of the loading pro- where Vgear1 is the maximum travel speed in first gear,
cess. That is, the operator will crowd the pile until Rimpullmax is the maximum rimpull force at full slip,
he/she senses the machine has slowed to some and is approximated as
nominal speed, at which he/she begins using the
lifting and tilting functions to fill the bucket in a Rimpullmax = g µd Mvehicle (34)

Fig. 11 Bucket tool geometry variables

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2240 M D Worley and V La Saponara

Moreover, H is the total horizontal force imparted to approximation desired, this detail is not considered in
the bucket by the terrain. In general, H is given by this study.
As mentioned previously, several preliminary stud-
ies were undertaken to determine a reasonable value
H = FV_P3 sin α + FH_P3 cos α (35)
for the parameter b in equation (29). After some itera-
tion, it was decided that setting b equal to one-half of
It is recognized that the maximum available rimpull the cutting edge thickness best matched the available
may vary with time – increasing due to vertical load- field data. Though not measured, the soil properties
ing on the bucket and/or an increasing mass due intrinsic to the collected pressure data are most sim-
to payload. However, staying with equation (34), this ilar to those of sandy soils. The soil properties given
could in fact easily lead to an estimate for rimpull that in reference [20] were used for comparison, with the
exceeds the power capability of the vehicle’s engine. dry sand and sandy loam soils most heavily influenc-
Since the powertrain response is omitted and a simpler ing the choice of b. Figure 12 shows the calculated

Fig. 12 Plots of calculated crowd loads (rimpull) and lift cylinder pressures for various soils using
b equal to one-half cutting edge thickness
Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2241

rimpull and lift cylinder pressures for crowding into area of the soil wedge – defined by A1 B1 C1 – represents
various soils. Figure 12(a) shows the two critical transi- the mass accumulated thus far. At some time later, the
tion parameters (the critical forward speed and critical bucket has moved through an arbitrary path, and the
penetration depth) that trigger the onset of Phase 2. new soil wedge has area defined by A2 B2 C2 . The mass
Figure 12(b) shows the estimated jump in pressure accumulated from the first state to the second – not
as Phase 2 begins, i.e. the supporting force from counting the current soil wedge – is represented by the
the pile vanishes. Note that the ‘clayey soil’ and the area A1 A2 B2 B1 . This area is calculated as
‘sandy loam4’ transitions are triggered by the critical
penetration distance rather than the critical speed. A1 A2 B2 B1 = E1 E2 B2 B1 + A1 E1 B1 − A2 E2 B2 (39)

2.3.2 Phase 2: bucket filling where the area E1 E2 B2 B1 is calculated using the trape-
zoid rule as
The motion of the cutting edge is due to a combination
of forward motion of vehicle, lift cylinder actuation, 1  
and tilt cylinder actuation. The transition from Phase 1 E1 E2 B2 B1 = χi di + di−1 (40)
2
to Phase 2 involves, by definition, the actuation of
either the tilt cylinder or the lift cylinder. Either case
results in loss of contact between the terrain and the This reasoning holds for any two wedges from one
underside of the cutting edge, eliminating the forces timestep to the next. Therefore the mass accumu-
Fsupport and Ca LBOC . The forces P, Ca LT , and F remain lated in sweeping the bucket from a configuration
active throughout Phase 2. Expressions for the loads at ψ1 = f (ρ1 , d1 , L1T ) to a configuration ψ2 = f (ρ2 , d2 , L2T )
point P3 take the form is approximated as

FV_P3 = F cos δ (36) 


N
Maccum = γ · w · Ai (41)
FH_P3 = P + F sin δ + Ca LT w (37) i=1
LT
MP3 = F cos δ (38)
2
In addition, the accumulating mass of the payload
must be tracked and added to the payload mass cen-
tre. For simplicity, this payload mass centre is assumed
to be constant throughout the digging process and is
defined per SAE J742. This is a reasonable assump-
T  LT  L T , as the mass of soil being
tion for 0.65L crit crit

pushed around the bucket curvature will tend to


accumulate above the soil wedge near to this point.
Referring to Fig. 13, accumulation of mass is calcu-
lated in the following way. At the end of Phase 1, the

Fig. 13 Geometry for cutting edge motion Fig. 14 Operator command template

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2242 M D Worley and V La Saponara

where N is the number of timesteps taken between of V requires taking an average of the velocities along
configurations, and where Ai , χi , and di are given below boundary AB at the soil wedge. The mass accumulated
 i−1 2  2 at each timestep is transferred to the payload mass
1   d cotρ i−1 − d i cot ρ i centre. At the end of Phase 2, all of the swept mass has
Ai = χi di − di−1 + accumulated at the payload centre, with the exception
2 2
(42) of the most recent soil wedge. Phase 2 ends when one
N P3   of the following events occurs:
χi = v i−1 · DP 1 + VFWDi−1 cos θ t (43)
 N P3 
di = di−1 + VFWDi−1 sin θ − v i−1 · DP 2 t (44) (a) the rake angle ρ drops below a nominal limit;
(b) the accumulated mass plus the current soil wedge
Note that the velocity component in brackets in mass exceeds a nominal limit;
equation (43) is not, in general, equal to the cutting (c) the volume of the accumulated mass plus the vol-
plane speed V used in equations (21) to (23). For gen- ume of the current soil wedge exceeds a nominal
eral planar motion of the bucket, the determination limit.

Fig. 15 (a) Lift cylinder pressures during control boom cycle (‘plot 1’); (b) tilt cylinder pressures
during control boom cycle (‘plot 2’). Field data is represented by solid lines

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2243

Fig. 16 (a) Lift cylinder pressures during control boom cycle (‘plot 3’); (b) tilt cylinder pressures
during control bucket cycle (‘plot 4’). Field data is represented by solid lines

2.3.3 Phase 3: rollback 3 OPERATOR INPUT

The tilt cylinder is actuated until the bucket is fully As mentioned previously, the operator action to end
rolled, which, depending on boom height, will occur crowding and begin filling the bucket is automated by
when either the tilt cylinder reaches full extension or achieving either the critical penetration depth or the
the mechanical roll stop is contacted. The mass of critical forward speed of the vehicle. During bucket fill-
the remaining soil wedge from Phase 2 is transferred ing, the operator response is somewhat autonomous
to the payload mass centre. Inherent in this approach as the only constraint is that his/her actions keep the
is the assumption that there is negligible soil cutting forward speed of the vehicle within a set bandwidth
force at this point in the digging process. around the nominal critical forward speed. The model

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2244 M D Worley and V La Saponara

is provided by a standard operator template that is fol- Based on the results, the simulation shows to be in
lowed if the velocity requirement is met. Otherwise, reasonable agreement with test results. There are pri-
deviations are made from the template to try and cor- marily two areas where the simulation and field data
rect the forward velocity. Figure 14 shows the template differ substantially. However, these can be explained
operator input for the digging portion of the short as follows:
work cycle.
1. Plot 1 (the top plot in Fig. 15) and plots 3 and
4 (the plots in Fig. 16) show different responses
4 COMPARISON WITH FIELD DATA AND over the dwell regions, where the initial cylinder
DISCUSSION actuation has reached its extent and the opera-
tor reverses the control lever to return to original
Two control simulations are performed and compared position (e.g. at 5.75–6.50 s for the boom cycle case
with field test data measured on an actual machine, and 3.0-3.75 s for the bucket cycle case). It should
to verify that the hydraulic cylinder and valve mod- be noted that in reality this will entail striking a
els are working together with the bucket mechanism mechanical stop – either inside the cylinder itself
dynamics. These control cases do not include terrain or between designated stop pads on the mecha-
loads from the previous section in order to isolate the nism links. This event was not modelled by the
hydraulics and bucket mechanism. simulation. Therefore, it is expected that the mea-
The first control case is referred to here as the boom sured data and simulation output will differ in these
cycle case. In this scenario, the lift cylinders are acti- regions.
vated to raise the boom to its full height, hold it at full 2. Referring to Plot 2 (the bottom plot in Fig. 15),
height for a certain period of time, then lower it back the lift cylinder pressures oscillate wildly about the
to its original starting position. The bucket tool has measured data. This makes sense considering that
no payload during the operation and pressures in the the hydraulic cylinder model provides little damp-
tilt cylinder are purely reactionary. The second con- ing when fluid is not flowing into or out of it (i.e.
trol case is referred to here as the bucket cycle case. In across the flow control valves). This is the case
this scenario, the tilt cylinder is activated to roll the here, where only the tilt cylinder is being ‘activated’
bucket tool from a full dump position to a fully rolled and the lift cylinders are reacting more or less like
position, hold it at fully rolled for a certain period of springs. In reality, viscous friction in the cylinder
time, then dump it back to its original starting posi- and mechanism joints, as well as the compliance
tion. Again, the mechanism has no payload and, in of the tires, provide damping to the system and
this case, the pressures in the lift cylinders are purely help mitigate these oscillations. The model appears
reactionary. under-damped without including some of these
In either case, cylinder actuation is controlled phenomena.
through changes in the control valve orifice area. Data
for this orifice area as a function of the control lever Next, simulations including the terrain forces on the
throw distance was used to generate equations that bucket were conducted. Table 1 provides an abridged
govern the exposed area as function of lever throw. summary of the simulation parameters used in the
Comparison of simulation output and measured val- correlation and dig studies. (Intellectual property pre-
ues is shown in Figs 15 and 16. vents detailed publication of mass properties data and
actual geometries.) Figure 17 shows a comparison of
the measured lift cylinder and tilt cylinder pressures to
Table 1 Parameters used in simulation
the simulation output.
Parameter Value The results show promise, but are certainly not as
accurate as one would like, even for simplified design
Maximum supply pressure, PS (N/m2 ) 24.8 E6
Tank (exhaust) pressure, PE (N/m2 ) 0 E6
schemes. The different results for both cylinders
Fluid bulk modulus, βF (N/m2 ) 2.0 E9 between the interval 3 and 4 s are believed to be due
Fluid density, ρhyd (kg/m3 ) 900 to a lack of modelling the compaction force called ‘f2 ’
Lift cylinder bore (mm) 125 by Hemami. Furthermore, it is believed that the dif-
Tilt cylinder bore (mm) 160
Dirt pile angle of repose, θ (◦ ) 37.5 ference in results for the tilt cylinder over the interval
Soil internal friction angle, φ (◦ ) 29 4–7 s is due to a lack of the extra soil weight existing on
Soil–tool friction angle, δ (◦ ) 17.5 the inclined slope that is not captured by the wedge
Bekker parameters
n 0.7 model, but was hoped to be ‘averaged’ out by keep-
KC (kN/m(n+1) ) 5.27 ing the Bekker sinkage force P active throughout the
Kφ (kN/m(n+1) ) 1515.04 loading process. The rise in tilt cylinder pressure over
Soil density, γ (T/m3 ) 1.4
Vehicle speed in low gear, Vgear1 (m/s) 2.0
the interval 7–9 s is thought to be due to the imme-
diate transfer of the soil wedge mass to the bucket

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2245

Fig. 17 Comparisons of the lift cylinder and tilt cylinder pressures (piston side) during digging
cycle. Field values are represented by solid lines

mass centre during the final rollback. In reality, the loads imparted to the bucket tool during the scoop-
mass transfer is more gradual (though, sometimes only ing/loading portion of the loader’s short work-cycle.
slightly), and the pile provides some support to the The equations of motion were developed using Kane’s
bucket while excess soil spills out of the bucket. method and programmed with the aid of a commercial
symbolic manipulator.
Test simulations isolating the hydraulic and mech-
anism models were run and compared with field data
5 CONCLUSIONS for proper correlation of control valve dynamics, a step
deemed necessary because of the extreme simplifi-
A subscale dynamic model of a unique front-end cation of the hydraulic system. Furthermore, several
loader bucket mechanism type was developed, that digging simulations were run to verify the capability of
included the hydraulic and mechanism dynamics, the model to predict the loads generated by the terrain
as well as simplified estimations for the terrain on the bucket. The current model assumptions appear

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2246 M D Worley and V La Saponara

to lack in the ability to capture these loads adequately of the International Symposium on Mine Mechanization
over the time history of the digging operation, most and Automation, Golden, CO, USA, June 1995, 12 pages.
notably during the bucket-filling phase. The model 9 Scheding, S., Dissanayake, G., Nebot, E., and Durrant-
does show promise in being able to predict peak oper- Whyte, H. Slip modeling and aided inertial navigation of
an LHD. In Proceedings of the IEEE International Con-
ation loads, seen most notably for the lift cylinder
ference on Robotics and Automation, Albuquerque, New
reaction loads. Future work and improvements should
Mexico, USA, 1997, pp. 1904–1909, vol. 3.
focus on improving the simulation of the bucket-filling 10 Available from http://www.autolev.com
phase. Also, similar simulations that vary the mech- 11 Kane, T. R. and Levinson, D. A. Dynamics, theory and
anism type, operator command, and soil properties applications, 1985 (McGraw-Hill, New York)
should be done to verify the robustness of the model. 12 Margolis, D. and Shim, T. Instability due to interacting
The field data required to do this is unfortunately not hydraulic and mechanical dynamics in backhoes. ASME
available at this time. J. Dyn. Syst. Meas. Control, 2003, 125, 497–504.
While this model’s accuracy cannot eclipse the accu- 13 Zhu, W. and Piedboeuf, J. Adaptive output force track-
racy provided by more detailed, full-scale models ing control of hydraulic cylinders with applications to
(e.g. those that utilize sophisticated operator logic to robot manipulators. ASME J. Syst. Meas. Control, 2005,
127, 206–217.
guide the vehicle), it is however promising for use
14 Singh, S. State of the art in automation of earthmoving.
as a preliminary performance analysis and for rapid
ASCE J. Aerosp. Eng., 1997, 10(4), 179–188.
characterization of design loads. 15 Hemami, A. Motion trajectory study in the scooping
operation of an LHD-loader. IEEE Trans. Ind. Appl., 1994,
30(5), 1333–1338.
ACKNOWLEDGEMENTS 16 Singh, S. Synthesis of tactical plans for robotic excava-
tion. PhD Thesis, The Robotics Institute, Carnegie Mellon
The authors thank Professor Don Margolis, Depart- University, Pittsburgh, PA, 1995.
ment of Mechanical and Aeronautical Engineering, 17 Ericsson, A. and Slattengren, J. Predicting digging
and cutting forces in granulated material. In the 15th
University of California, Davis, for his feedback and
European ADAMS Users’ Conference, Rome, Italy,
review of this paper. 15–17 November 2000, available from http://www.
mscsoftware.com/support/library/conf/adams/euro/
2000/Volvo_Predicting_Digging.pdf.
REFERENCES 18 Bekker, M. G. Introduction to terrain-vehicle systems,
1969 (University of Michigan Press, Michigan).
1 Filla, R. Operator and machine models for dynamic 19 Marshall, J. A. Towards autonomous excavation of
simulation of construction machinery. Licentiate The- fragmented rock: experiments, modelling, identification,
sis, Department of Mechanical Engineering, Linkopings and control. MS Thesis, Queen’s University, Kingston,
Universitet, Linkoping, Sweden, 16 September 2005. Ontario, Canada, 2001.
2 Terzaghi, K. Theoretical soil mechanics, 1947 (Wiley, 20 Wong, J. Y. Theory of ground vehicles, 1979 (John Wiley &
New York). Sons, New York).
3 McKyes, E. Soil cutting and tillage, 1985 (Elsevier Press,
Amsterdam).
4 Filla, R., Ericsson, A., and Palmberg, J.-O. Dynamic sim-
ulation of construction machinery: towards an operator
model. In the IFPE 2005 Technical Conference, Las Vegas, APPENDIX
NV, USA, 16–18 March 2005, pp. 429–438.
5 Filla, R. An event-driven operator model for dynamic Notation
simulation of construction machinery. In the Ninth
Scandinavian International Conference on Fluid Power, abucket_cg acceleration of the bucket centre
Linkoping, Sweden, 1–3 June 2005, available from of gravity
http://arxiv.org/ftp/cs/papers/0506/0506033.pdf. apayload_cg acceleration of the centre of
6 Filla, R. and Palmberg, J.-O. Using dynamic simula- gravity of accumulated soil
tion in the development of construction machinery. In APIS area of piston-side cross-section
the Eighth Scandinavian International Conference on of hydraulic cylinder
Fluid Power, Tampere, Finland, 7–9 May 2003, vol. 1, AROD area of rod-side cross-section of
pp. 651–667.
hydraulic cylinder
7 Hemami, A. and Daneshmend, L. Force analysis for
automation of the loading operation in an LHD loader.
AVALVE area of hydraulic control valve
In Proceedings of the 1992 IEEE International Confer- orifice
ence on Robotics and Automation, Nice, France, 1992, b effective width of penetrating
pp. 645–650. surface
8 Sarata, S., Sato, K., and Yuta, S. Motion control system (BKT1 , BKT2 ) reference frame local to bucket
for autonomous wheel loader operation. In Proceedings (BKTCG1, location of centre of gravity

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2247

BKTCG2) of bucket in its local reference GDLKCG2) gravity of guide link in its local
frame reference frame
(BKTLK1 , reference frame local to H total horizontal load on bucket
BKTLK2 ) bucket link cutting edge
(BKTLKCG1, location of centre of J1 fixed pivot joint of body boom
BKTLKCG2) gravity of bucket link in its local J2 fixed pivot joint of tilt cylinder
reference frame J3 fixed pivot joint of lift cylinder
(BLCRK1 , reference frame local to J4 fixed pivot joint of ‘Y-link’
BLCRK2 ) bellcrank J5 pivot joint between lift cylinder
(BLCRKCG1, location of centre of and boom
BLCRKG2) gravity of bellcrank in its local J6 pivot joint between bellcrank and
reference frame bucket link
(BOC1 , BOC2 ) reference frame local to bucket J7 pivot joint between bellcrank and
bolt-on cutting edge ‘Y-link’
(BOCHORIZ , location of load application J8 pivot joint between bellcrank and
BOCVERT ) point (P3) on the bucket in the tilt cylinder
local bolt-on cutting edge J9 pivot joint between guide link and
reference frame bucket link
(BOOM1 , reference frame local J10 pivot joint between bucket link
BOOM2 ) to boom and bucket
(BOOMCG1, location of centre of J11 pivot joint between guide link and
BOOMCG2) gravity of boom in its local boom
reference frame J12 pivot joint between boom and
C cohesion factor for the soil bucket
C1 , C2 correlation constants for flow kC cohesive modulus of
across control valve orifice deformation – empirical factor
Ca cohesion factor between soil and used in Bekker’s load-sinkage
cutting tool formula
d projection normal to dirt pile free kφ frictional modulus of
boundary of bucket cutting edge deformation – empirical factor
penetration depth used in Bekker’s load-sinkage
N
d()/dt derivative with respect to time in formula
reference system N L1A, L1B, L1C lengths along boom
(DP1 , DP2 ) reference frame local to the dirt L2 length between joints J7 and J8
pile free boundary L3 length between joints J6 and J7
F net normal and friction force L4 length between joint J5 and point
along boundary between cutting P1
tool and soil wedge L5 length between joint J11 and
FA inertial force generated in point P2
accelerating soil mass L6 length between joints J4 and J7
Fcyl net force generated by hydraulic (i.e. length of ‘Y-link’)
pressure in cylinder L7 length between joints J6 and J9
FH_P3 component of total digging load L8 length between joints J9 and J10
acting on the bucket cutting edge, L9 length between joints J9 and J11
resolved parallel to the cutting (i.e. length of guide link)
edge motion LBOC length of boundary between soil
Fsupport net supporting force of dirt pile wedge and bolt-on cutting edge
on underside of cutting tool (i.e. (LCYL1 , LCYL2 ) reference frame local to lift
bucket) cylinder
FV _P3 component of total digging load LCYLPISCG1 distance to centre of gravity of
acting on the bucket cutting edge, barrel section of lift cylinder in its
resolved normal to the cutting local reference frame
edge motion LCYLRODCG1 distance to centre of gravity of rod
g gravitational constant section of lift cylinder in its local
(GDLK1 , reference frame local to reference frame
GDLK2 ) guide link LF length of boundary between soil
(GDLKCG1, location of centre of wedge and dirt pile

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


2248 M D Worley and V La Saponara

LLCYL overall length of the lift cylinder TCYLPISCG1 distance to centre of gravity of
(joint-to-joint) barrel section of tilt cylinder in its
LTCYL overall length of the tilt cylinder local reference frame
(joint-to-joint) TCYLRODCG1 distance to centre of gravity of rod
LT length of boundary between section of tilt cylinder in its local
cutting tool and soil wedge reference frame
mbucket mass of the bucket U1 generalized speed set as the
mpayload mass of the accumulated soil velocity of lift cylinder rod
MP3 net moment acting on the bucket U2 generalized speed set as the
cutting edge velocity of tilt cylinder rod
Mvehicle total mass of the vehicle U3 generalized speed set as the
n exponent of angular velocity of lift cylinder
deformation – empirical factor U4 generalized speed set as the
used in Bekker’s load-sinkage angular velocity of tilt cylinder
formula U5 generalized speed set as the
(N1 , N2 ) inertial reference frame angular velocity of boom
Na Terzaghi factor for soil cutting U6 generalized speed set as the
force (acceleration term) angular velocity of bellcrank
NC Terzaghi factor for soil cutting U7 generalized speed set as the
force (soil cohesion term) angular velocity of bucket link
NCa Terzaghi factor for soil cutting U8 generalized speed set as the
force (soil–tool cohesion term) angular velocity of bucket
Nq Terzaghi factor for soil cutting U9 generalized speed set as the
force (surcharge term) angular velocity of guide link
Nγ Terzaghi factor for soil cutting U10 generalized speed set as the
force (soil shear term) angular velocity of ‘Y-link’
p penetration pressure on bucket vLCYL_ROD linear velocity of lift cylinder
cutting edge rod
P penetration force on bucket vROD linear velocity of cylinder rod
cutting edge (general case)
P3 load application point on the vTCYL_ROD linear velocity of tilt cylinder rod
bucket V average speed of cutting tool
PE hydraulic system exhaust (tank) V relative speed of cut soil mass
pressure VFWD forward speed of vehicle during
PPIS hydraulic pressure on piston side penetration
of cylinder Vgear1 maximum forward speed of
PROD hydraulic pressure on rod side of vehicle in first gear
cylinder VolPIS hydraulic fluid volume in
·
P PIS change in piston-side hydraulic piston-side chamber of cylinder
pressure with respect to time VolROD hydraulic fluid volume in rod-side
·
P ROD change in rod-side hydraulic chamber of cylinder
pressure with respect to time w width of the bucket cutting edge
PS hydraulic system supply (pump) Wbucket weight of the bucket
pressure Wpayload weight of the accumulated mass
q surcharge pressure (soil) in the bucket
QPIS flowrate on piston side of cylinder Wwedge weight of the soil wedge
QROD flowrate on rod side of cylinder (X 1, Y 1) location of joint J1 in inertial
R net normal and friction force reference frame
along boundary between soil (X 2, Y 2) location of joint J2 in inertial
wedge and dirt pile reference frame
Rimpullmax maximum push force that vehicle (X 3, Y 3) location of joint J3 in inertial
can develop (either powertrain or reference frame
traction limited) (X 4, Y 4) location of joint J4 in inertial
t time reference frame
(TCYL1 , TCYL2 ) reference frame local to tilt (YLINK1 , reference frame local to
cylinder YLINK2 ) ‘Y-link’

Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science JMES688 © IMechE 2008

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016


Dynamic model for front-end loader design 2249

(YLINKCG1, location of centre of γBKTLK fixed angle of bucket link


YLINKCG2) gravity of ‘Y-link’in its local geometry
reference frame γBLCRK fixed angle of bellcrank geometry
z penetration depth of cutting δ soil–tool friction angle
tool θ angle of repose of soil
centre of gravity µd dynamic coefficient of friction
• geometrical point of interest in between vehicle tires and ground
figures defining kinematics, also ρ rake angle of bucket during soil
indicates connection point in cutting
hydraulic schematic ρhyd density of hydraulic system fluid
figures φ soil internal friction angle
 a planar pivot joint in figures χ distance along dirt pile free
defining kinematics, also boundary travelled by cutting tool
indicates a ball check valve in over a given instant of time
hydraulic schematic figures χ distance along dirt pile free
boundary travelled by cut mass of
α inclination angle of bucket soil over instant of time
cutting edge relative to ground BKTLK angular velocity of bucket link
(global horizontal). This is BKT angular velocity of bucket
equivalent to orientation angle of BLCRK angular velocity of bellcrank
the bolt-on cutting edge BOOM angular velocity of boom
coordinate system reference, φBOC GDLK angular velocity of guide link
βF bulk modulus of hydraulic fluid LCYL angular velocity of lift cylinder
β break angle of soil wedge TCYL angular velocity of tilt cylinder
γ soil density YLINK angular velocity of ‘Y-link’

JMES688 © IMechE 2008 Proc. IMechE Vol. 222 Part C: J. Mechanical Engineering Science

Downloaded from pic.sagepub.com at PENNSYLVANIA STATE UNIV on May 11, 2016

You might also like