Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

This article was downloaded by: [University of Sydney]

On: 05 September 2014, At: 11:56


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Vehicle System Dynamics: International


Journal of Vehicle Mechanics and
Mobility
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/nvsd20

Tractor–semitrailer model for vehicles


carrying liquids
a a
Mohammad Biglarbegian & Jean W. Zu
a
Department of Mechanical and Industrial Engineering ,
University of Toronto , 5 King’s College Road, Toronto, Ontario,
Canada , M5S 3G8
Published online: 17 Feb 2007.

To cite this article: Mohammad Biglarbegian & Jean W. Zu (2006) Tractor–semitrailer model for
vehicles carrying liquids, Vehicle System Dynamics: International Journal of Vehicle Mechanics and
Mobility, 44:11, 871-885, DOI: 10.1080/00423110600737072

To link to this article: http://dx.doi.org/10.1080/00423110600737072

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or
arising out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Vehicle System Dynamics
Vol. 44, No. 11, November 2006, 871–885

Tractor–semitrailer model for vehicles carrying liquids


MOHAMMAD BIGLARBEGIAN* and JEAN W. ZU
Department of Mechanical and Industrial Engineering, University of Toronto, 5 King’s College Road,
Toronto, Ontario, Canada M5S 3G8
Downloaded by [University of Sydney] at 11:56 05 September 2014

This article presents a model for solving solid–fluid interactions in vehicles carrying liquids. A tractor–
semitrailer model is developed by incorporating suspension systems and tire dynamics. Owing to
the solid–fluid interaction, equations of motion for the vehicle system are coupled. To simplify the
complicated solution procedure, the coupled equations are solved separately using two different codes.
Each code is analyzed separately; but as the parameters of the two codes depend on each other, the codes
must be connected at the end of each time step. To determine the dynamic behavior of the system,
different braking moments are applied. As the braking moments increase, braking time decreases.
However, it turns out that increasing the braking moment to more than a certain level produces no
significant results. It is also shown that vehicles carrying fluids need a greater amount of braking
moments in comparison to vehicles carrying solids during braking. In addition, as the level of the fluid
inside the tanker increases, from one-third to two-third of the tanker’s volume, the sloshing forces
applied to the tanker’s walls increase. It was also concluded that the strategy used in this article to
solve for the solid–fluid interaction by incorporating vehicle dynamic effects represents an effective
method for determining the dynamic behavior of vehicles carrying fluids in other critical maneuvers.

Keywords: Solid–fluid interaction; Sloshing; Tractor–semitrailer; Vehicle dynamics; Tire model;


Lagrange’s equations

1. Introduction

Owing to the catastrophic nature of accidents, the handling of vehicles containing liquid cargo
is of great importance. Specifically, for different critical situations such as braking or turning,
determining the effect of sloshing on dynamic behavior of the vehicle is necessary.
Many studies have been performed to capture the behavior of systems containing solids and
fluids [1–8]. With regards to braking maneuver models, Ranganathan et al. [2] developed a
mass–spring–damper model to investigate the dynamic characteristics of partially filled liquid
tanks during typical straight-line braking maneuvers. Pitch plane model was also adopted to
investigate the straight-line braking performance of partially filled tanks subject to constant
decelerations by Rakheja and Wang [3] and Ranganathan and Ying [4]. In addition to braking,
another important factor is determining the rollover stability of vehicles carrying liquid cargo
during critical turning. Rakheja et al. [6] proposed a roll plane model to study the steady-state
load shift and static roll stability of tank vehicles. Also, Papov et al. [9] proposed a roll moment
model to determine the rollover threshold of tank trucks with different cross-sections: circular,

*Corresponding author. Email: mbiglar@mie.utoronto.ca

Vehicle System Dynamics


ISSN 0042-3114 print/ISSN 1744-5159 online © 2006 Taylor & Francis
http://www.tandf.co.uk/journals
DOI: 10.1080/00423110600737072
872 M. Biglarbegian and J. W. Zu

elliptic, and modified oval. Similar to these models, frequency shape backstepping sliding
mode was proposed by Acarman and Ozguner [10] to stabilize the sloshing effects of moving
cargo.
In order to determine the effect of sloshing and also predict the rollover stability of vehicles
carrying fluids, accurate simulations must be made. Sankar et al. [11] solved the problem
of coupled sloshing liquids and rigid body structures by using the interaction of CFD and
multibody system solver methods. Nevertheless, this method is highly computational and thus
requires a significant amount of computational effort in large scale systems. Rumold [12]
used the multigrid algorithm to solve the discretized Navier–Stokes equation to analyze the
interactions of sloshing liquids with rigid body systems in which a two-dimensional model
was considered to adequately characterize the dynamic behavior of the system. While he
solved the coupled equations of solid and fluid system, a simplified two-dimensional model
Downloaded by [University of Sydney] at 11:56 05 September 2014

was adopted for vehicle body system. Most recently, a computer code was developed in which
a simplified three-dimensional dynamic model for vehicle was used [13]. A free surface code
in this program was developed, which could simulate the sloshing of liquid inside a tanker.
This work is well capable of determining the behavior of the fluid on braking subject to the
information given from the vehicle system. However, a simplified three-dimensional model
was adopted for developing the vehicle system. In this work, suspension system and wheel
dynamics were not considered to develop the model.
With respect to the vehicle system itself, different models of single tractors and tractor–
semitrailers were developed to study the dynamic behavior of heavy vehicles during different
maneuvers [14–25]. Mikulcik [20] developed a dynamic model of a tractor–semitrailer to study
the rollover phenomenon. In this model, unsprung masses were treated as rigid bodies and
hence the influence of the tires on the vehicle roll motion was neglected. Ellis [18] developed
a tractor–semitrailer model with six degrees of freedom, in which a combined stiffness of
individual suspensions and tires was used to study the effect of the tires on roll motion. Chen
and Tomizuka [29] adopted a five degree-of-freedom tractor–semitrailer model by neglecting
the tractor pitch and bouncing motions for the purpose of lateral control of vehicles. Hyun [23]
adopted a 14 degree-of-freedom vehicle model to develop a predictive modeling and active
control of rollover in heavy vehicles.
Regarding vehicles carrying liquids, although numerous fluid models have been developed
to capture the dynamic behavior of liquids inside tankers, few models have been developed
to address the challenges faced by rigid body vehicle systems. As the vehicle system consists
of different parts – namely, suspension and tire elements – the effect of these subsystems must
be considered within the context of a vehicle model. In addition, by using tire models, Magic
Formula [16, 17], more accurate model for tires can be developed. In this article, a complete
dynamic model, namely tractor–semitrailer, to determine the dynamic behavior of vehicles
subject to liquid slosh loads during braking is presented. The main contributions of this article
lie in: solving fluid–solid interactions and developing a complete tractor–semitrailer model of
vehicle by considering suspension system and tire dynamics.

2. Dynamic model

To solve the problem of fluid–solid interaction, the vehicle and the fluid inside the tanker are
divided into two subsystems: the fluid subsystem and the rigid subsystem. The latter is referred
to as the dynamic subsystem. This division enables us to model the two systems separately.
As these two subsystems are related to each other, the two subsystems should be connected.
This is accomplished at the end of each procedure. In this article, first the dynamic model
Tractor–semitrailer model for vehicles 873

is presented. Subsequently, a procedure for fluid–solid interactions will be introduced and


discussed.

2.1 Vehicle body model

A tractor–semitrailer model is depicted in figure 1. The model has three degrees of freedom:
translational, roll, and pitch motion. Other degrees of freedom have been neglected because the
aforementioned degrees of freedom are well capable of capturing the whole dynamic behavior
of the vehicle.
Five coordinate systems for the tractor–semitrailer model are considered. As shown in
figures 1 and 2, the first coordinate system is the inertial coordinate system (xyz)n which is
fixed to the ground, the second is mounted to the unsprung mass (xyz)u (figure 2), the third
Downloaded by [University of Sydney] at 11:56 05 September 2014

is fixed to the sprung mass (xyz)ϕ at the rollover center, the fourth is fixed at the pitch center
(xyz)θ and finally at the intersection of tractor and semi-trailer, where the relative motion
between tractor and semitrailer happens (xyz)β .
The forces applied to the whole vehicle system are fluid forces, fluid moments, and tire
forces. Fluid forces and moments are assumed to be applied at the center of the tanker and

Figure 1. Kinematics diagram of the tractor–semitrailer model.

Figure 2. Rear view: suspension system.


874 M. Biglarbegian and J. W. Zu

Figure 3. Forces and moments applied to the tractor–semitrailer.


Downloaded by [University of Sydney] at 11:56 05 September 2014

they are known from fluid dynamic analysis and shown in figure 3. Tire forces are longitudinal
and lateral forces applied from the ground.

2.2 Equations of motion of vehicle body

Lagrange approach appears to be more convenient for the purposes of deriving equations
of motion. The principles in this method are based on an overview of the system and its
mechanical energy (kinetic and potential). In contrast, the Newton equations of motion are
time derivatives of momentum principle [28]. In Newton–Euler formulation, constraint forces
and moments must be taken into account when more than one body is considered. Doing so
increases the solutions’ level of complexity.
The standard form of Lagrange’s equation for a system with n independent generalized
coordinates can be written as
 
d ∂L ∂L
− = Qi (i = 1, 2, . . . , n) (1)
dt ∂ q̇i ∂qi
where L is the Lagrangian function, qi the generalized coordinate, and Qi the generalized
force. The Lagrangian, L, is defined as

L=T −V (2)

where T and V represent kinetic and potential energy of the entire system. The generalized
forces can be obtained as

n
∂xj
Qi = Fj (j = 1, . . . , n) (3)
i=1
∂qi

where Fi is the applied force at the ith coordinate xi .


The kinetic and potential energies are expressed as
1  2  1  
T = m1 vG1x + vG 2
+ v 2
G1z + m 2 v 2
G2x + v 2
G2y + v 2
G2z
2 1y
2
1  (4)
+ 2
Ixx1 ω1x + Iyy1 ω1y
2
+ Izz1 ω1z
2
+ Ixx2 ω2x
2
+ Iyy2 ω2y2
+ Izz2 ω2z
2
2
− (Ixz1 ω1x ω1z + Iyz2 ω2y ω2z )
V = −(m1 gh1 + m2 gh2 ) = −(m1 grG1 (3) + m2 grG2 (3)) (5)
Tractor–semitrailer model for vehicles 875

In equations (4) and (5), m, I , ω, g, h, r denote mass, moment of inertia, angular velocity,
gravitational acceleration, center of mass height, and center of mass position, respectively.
Indices 1 and 2 refer to tractor and trailer accordingly.
In order to calculate the kinetic and potential energies of the system, the relationship between
the coordinate systems must be known. In general, for two arbitrary coordinate systems,
elements of transformation matrices are the cosine angles between coordinate axes. For each
manoeuvre, corresponding transformation matrices will be set up and defined as follows:
   
1 0 0 1 0 0
n
Tu = 0 1 0 u
Tr = 0 cos ϕ − sin ϕ 
0 0 1 0 sin ϕ cos ϕ
    (6)
cos θ 0 sin θ 1 0 0
Downloaded by [University of Sydney] at 11:56 05 September 2014

r
Tp =  0 1 0  p TG = 0 1 0
− sin θ 0 cos θ 0 0 1

Velocity of the mass center of tractor can be obtained using the following relationships:

VG1 = Vop + ω
 p × rpG + Vrel/pG1 (7)
Vop = Vor + ω
 r × rrp + Vrel/rp (8)
Vor = Vou + ω
 u × rur + Vrel/ur (9)

where VG1 , Vop , Vor , Vou are the velocities of points G1 , p, r, u respectively; ωp denotes the
angular velocity of the tractor (sprung mass) relative to the sprung mass without considering
roll motion, ωr denotes sprung mass relative to unsprung mass angular velocity, and ωu denotes
the usprung mass angular velocity. Vrel/pG1 , Vrel/rp , Vrel/ur indicate relative velocities among
points pG1 , rp, u correspondingly. In equations (7) to (9), Vrel terms are zero as all points
belong to one rigid-body system.
Combining equations (7) to (9), we get

VG1 = Vou + ω
 u × rur + ω
 r × rrp + ω
 p × rpG1 (10)

where VG1 is the velocity of the tractor mass center.

VG2 = Vou + ω
 u × rur + ω
 r × rrp + ω
 p × rpt + ω
 t × rtG2 (11)

where VG2 is the velocity of the semi-trailer mass center.


Position of the tractor mass center is given by
     
xn rpx pG1x
rG1 = n Tu yn  + n Tr rpy  + n Tp pG1y  (12)
z0 rpz r pG1z p

Position of the trailer mass center is given by


       
xn 0 rpx pG2x
rG2 = yn  + n Tu  0  + n Tr rpy  + n Tp pG2y  (13)
0 z0 rpz pG2z
876 M. Biglarbegian and J. W. Zu

Using equations (4) to (13), kinetic and potential energies will be given as

1 1
T = (I1y + I2y )θ̇ 2 + m1 (vy2 + (pG1z θ̇ + vx − pG1x θ̇θ)2
2 2
1 (14)
+ (−pG1x θ̇ − pG1z θ̇ θ )2 ) + m2 (vy2 + (pG2z θ̇ + vx − pG2x θ̇θ)2
2
+ (−pG2x θ̇ − pG2z θ̇ θ )2 )
V = −(m1 gh1 + m2 gh2 ) = −(m1 grG1 (3) + m2 grG2 (3)) (15)

Generalized forces are as follows:


Downloaded by [University of Sydney] at 11:56 05 September 2014


n
∂rxti 
n
∂ryti 
n
∂rpti ∂rrtfx ∂rrtfy
Qθ = Fx n i + Fy n i + Fp n i + F lx + Fly
i=1
∂θ i=1
∂θ i=1
∂θ ∂θ ∂θ

∂rrtfz n
∂rxti 
n
∂ryti 
n
∂rpti ∂rrtfx
+ Flz + Mlz = Fx i + Fy 1 + Fp n i + Flx
∂θ i=1
∂θ i=1
∂θ i=1
∂θ ∂θ
∂rrtfy ∂rrtfz
+ Fly + Fl z+ Mlz = (−l − l1 )Fp1 + (−l − l1 )Fp2 + (−l − l2 )Fp3
∂θ ∂θ
+ (−l + l2 )Fp4 + (−l − l3 )Fp5 + (−l + l3 )Fp6 + Flx (pG2z cos θ − pG2x sin θ)
+ F lz (−pG2x cos θ − pG2z sin θ ) + Mly
(16)

Applying Lagrange’s equations, equations of motion are derived and shown as follows:

m1 θ̈ pG1z + m1 θ̇ 2 pG1x + m1 θ θ̈ pG1x − m2 θ̈pG2z + m2 θ̇ 2 pG1x + m2 θ θ̈ pG2x


ẍ = (17)
m1 + m 2
 
−m1 ẍpG1z + m1 ẍθ pG1x − m2 θ̇ 2 pG22z θ − m2 θ̈ pG2z − m1 θ̇ 2 pG21x θ − m2 θ̇ 2 pG22x θ
−m1 θ̇ 2 pG21z θ − m2 ẍpG2z + m2 ẍθ pG2x − m2 gpG2x − m1 gpG1x
θ̈ =
m1 pG21x (θ 2 + 1) + m1 pG21z (θ 2 + 1) + m2 pG22x (θ 2 + 1) + m2 pG22z (θ 2 + 1)
+ Qθ (18)

where
 
−(p1 + p2 )(lf + la ) + (p3 + p4 )(lr − la ) + (p5 + p6 )(lb + lt ) + MLy
+ FLx (pG2z cos θ − pG2x sin θ ) + FLz (−pG2x cos θ − pG2z sin θ)
Qθ = (19)
m1 pG21x (θ 2 + 1) + m1 pG21z + m2 pG22x (θ 2 + 1) + m2 pG22z + I1y

2.3 Tire model

To obtain the forces applied from the ground to the tires, a free body diagram of the tire
is considered. The free body diagram is shown in figure 4. By writing the Euler’s equation
(equation (7)) with respect to the mass center of the wheel, the expression for angular accel-
eration of the wheel and also forces and moments applied to the tire is obtained. With regards
to the tire model, Magic Formula is used to obtain the longitudinal forces applied from the
Tractor–semitrailer model for vehicles 877
Downloaded by [University of Sydney] at 11:56 05 September 2014

Figure 4. Free body diagram of a tire.

ground to the tires.


M t − Mb − Mr − Fx
ω̇ = (20)
Iw
where Mt , Mb , Mr , Fx are traction, braking, rolling-resistant moments, and tractive force,
respectively. Rolling-resistance moment can be obtained as follows [15]:
Mr = µr |Fz | · r (21)
In recent years, an empirical method for characterizing tire behavior known as the Magic For-
mula has been developed and used in vehicle handling simulations [27]. The Magic Formula,
in its basic form, can be used to fit experimental tire data for characterizing the relationships
between the cornering force and slip angle and braking effort and longitudinal slip [26, 27].
y(x) = D sin{C arctan[Bx − E(Bx − arctan Bx)]} (22)
Y (X) = y(x) + Sv , x = X + Sh (23)
where Y (x) represent cornering force, self-aligning torque, or braking effort, and X denotes
slip angle or skid. Coefficient B is called stiffness factor, C shape factor, D peak factor, and E
curvature factor. Sh and Sv are the horizontal shift and vertical shift, respectively. Description
of each parameter and the associated numerical value can be found in ref. [29].

2.4 Suspension model

The suspension system usually consists of spring and dampers. Assuming linear springs and
dampers, suspension forces can be obtained using the following relationship:

Fsuspension = ˙ i ) (i = 1, 2, . . . , 6)
(ki i + ci  (24)
i

where ki is the constant of each spring, i the vertical deflection of each spring, ci the damping
coefficient of each spring, and ˙ i the velocity of vertical deflection of each damper. For each
maneuver, i and  ˙ i are defined as follows:

1 = −(la + lf )θ 2 = −(la + lf )θ
3 = (lr − la )θ 4 = (lr − la )θ (25)
5 = (lb + lt )θ 6 = (lb − lt )θ
878 M. Biglarbegian and J. W. Zu
Downloaded by [University of Sydney] at 11:56 05 September 2014

Figure 5. Flowchart of vehicle model and its interaction with fluid.

2.5 Dynamic flowchart

After deriving the equations of motion, now we can obtain a dynamic model for the vehicle
system. The flowchart in figure 5 specifically depicts the dynamic system and its subsystems.
The interaction of solid system with the fluid system is also shown. Inputs to the dynamic
model from the fluid model are force and moments FL , ML . These forces and moments are
assumed to be applied at the mass center of semitrailer of the two-body and mass center of
the tanker for the single tractor models. The system’s characteristics, constants, and geometric
parameters of the two systems are saved in a file and can be accessed during the simulation.
Initial conditions such as initial velocity and position will be applied to the dynamic system
as well. Using the initial conditions, equations of dynamic system can be solved knowing the
fluid forces, fluid moments, and also forces applied from the tires. The outputs of dynamic
model are linear and angular velocities V and ω. Linear velocity along with traction and
braking moments Tt , Mb are used to solve the wheel dynamics’ equations. Angular velocities
of each wheel ωi and also linear velocity obtained from the dynamic system V are used to
find the longitudinal slip of each tire λi . Thereafter, longitudinal slip along with slip angles si
are entered as inputs to tire model, which provide longitudinal and lateral forces Fx , Fy .
Input variables to the fluid model are angular velocity ω, angular acceleration ω̇ = α, trans-
lational acceleration R̈ = a, and transformation matrix S, which are obtained from dynamic
system analysis.

3. Fluid model

In this section, the fluid system and its governing equations are reviewed. The equations of
fluid system are conservation of mass and momentum. These equations are written as
Continuity:
∇ · V = 0 (26)
Momentum:

∂ V 1 1 1
+ ∇ · (V V ) = − ∇p + ∇ · µ(∇ V + (∇ V )T ) + g + Fb (27)
∂t ρ ρ ρ

where V is the velocity vector of fluid relative to the global system of coordinates, p the
pressure, and ρ and µ the density and kinematic viscosity of the liquid, respectively. g is the
Tractor–semitrailer model for vehicles 879

Figure 6. Global and fluid frames.

gravitational acceleration and Fb represents any body force (per unit volume) acting on the
Downloaded by [University of Sydney] at 11:56 05 September 2014

fluid. In order to track the free surface of the fluid, volume of fluid method (VOF) is used [30].
In this method, the volume tracking is done by introducing a scalar function f whose value is
one within the liquid phase and zero at other places. At the surface cells of the fluid, f has a
value between one and zero. The function f satisfies the advection equation
∂f
+ (V · ∇)f = 0 (28)
∂t
Solving the above equations is complicated considering the motion of the container. The fluid
equations must be written in the fluid reference frame (fixed to the container) to be independent
of the mechanical system at each time step. Therefore, the aforementioned equations must
be derived by considering the relationship between the global and fluid coordinate systems.
Figure 6 shows the two reference frames: global and fluid frames. It illustrates how the position
of a fluid particle can be expressed in the two frames.
After solving the fluid equations, fluid forces and moments applied to the rigid body can be
obtained. The fluid forces and moments are denoted by FL and ML , respectively. FL can be
calculated by integrating the following integral which is the integration of stress over the wet
surface of the container wall Sw :
FL = τ ds (29)
Sw

Similarly, the resulting fluid moment is given by

ML = rs × τ ds (30)
Sw

where rs is the position of the surface element with respect to the fluid frame. τ , strain, is given
by the following equation:

τ = [−p n̂ + µ(∇ V + (∇ V )T )
n] (31)

where p is the fluid pressure, n is normal to the surface element, and V is the fluid velocity
with respect to the global system.

4. Solid–liquid interactions

As mentioned in the introduction, vehicles carrying fluids have two subsystems: solid and
fluid. Owing to the fluid–solid interactions, the equations of motion for the vehicle system are
880 M. Biglarbegian and J. W. Zu

Figure 7. Fluid–solid interaction.


Downloaded by [University of Sydney] at 11:56 05 September 2014

coupled, and thus, solving the coupled equations is extremely complicated. In this section,
an effective strategy is presented for solving the coupled subsystems of fluid and solid. This
strategy involves decoupling the equations of motion by dividing the vehicle system into solid
and fluid subsystems. By doing so, the equations of each subsystem can be solved separately.
The flowchart in figure 7 depicts how the fluid and solid subsystems are connected. It
also shows how the information is exchanged among different parts of the system: inputs
(constants), solid subsystem, and the fluid subsystem. The input consists of systems such as
mass, inertias, dimensions, and density. The two subsystems use the information embedded
in the input and interact with each other. The solid system inputs fluid forces and moments
and outputs angular velocity, angular acceleration, linear acceleration, and transformation
matrices (kinematic parameters). In a similar fashion, the fluid subsystem inputs the kinematic
parameters and outputs fluid forces and moments. To understand how the subsystems interact
with each other, let us assume that at a specific time, t, input parameters to one subsystem
are known. For example, if that subsystem is the fluid, it is assumed that angular velocity and
translational and angular accelerations of the tanker are known (these parameters are obtained
on the basis of dynamic subsystem analysis). As the input parameters to the subsystem are
known, that subsystem can be analyzed independently and thus, output parameters can be
obtained. These outputs are now the input parameters for the other subsystem. For the given
example, fluid forces and moments are the inputs to the solid subsystem. Now, given the inputs
to the other subsystem, a very small increase is given to time t. That subsystem can now be
analyzed and outputs can be obtained. These outputs are inputs to the other subsystem, and so
on. This procedure continues until the desired time is reached. To ensure that the solutions do
not diverge, t must be very small, and t is found using iteration methods. In this article,
the coupled subsystems were solved several times for different time steps and it was apparent
that t value less than 0.003 s will converge to solutions.

5. Simulation and results

Two codes in the form of two different subroutines were developed to solve the equations
of the whole system, one for the rigid system and one for the fluid system. A code for the
dynamic system is developed in this article and a code for the fluid system is taken from an
existing code. System’s characteristics such as mass, moment of inertia, masses and moments
inertia of the wheels, and dimensions of the system are shown in table 1. The aforementioned
codes were written as two Fortan subroutines and run for different braking moments. Figures 8
and 9 show position, velocity, and acceleration of single tractor model for different braking
moments of 3000, 5000, and 6000 N m. Simulations were carried out for a time period of 5 s.
Tractor–semitrailer model for vehicles 881

Table 1. System characteristics of the tractor–semitrailer model.

Parameter Value Parameter Value

m1 (kg) 5.450E+003 xt (m) 4.000


m2 (kg) 1.300E+004 yt (m) 1.520
I1x (kg m2 ) 2.456E+003 zt (m) 1.760
I1y (kg m2 ) 3.224E+003 tG2y (m) 0.000
I1z (kg m2 ) 2.866E+003 tG2z (m) −0.880
I1xy (kg m2 ) 0 Kf 1 (N/m) 2.50E+005
I1xz (kg m2 ) 0 Kf 2 (N/m) 3.36E+0010
I1yz (kg m2 ) 0 Kr1 (N/m) 6.87E+005
I2x (kg m2 ) 5.859E+003 Kr2 (N/m) 1.05E+0011
I2y (kg m2 ) 2.069E+004 Kt1 (N/m) 1.55E+006
I2z (kg m2 ) 1.984E+004 Kt2 (N/m) 1.92E+0012
I2xy (kg m2 ) 0 Cf (N s/m) 9080.000
Downloaded by [University of Sydney] at 11:56 05 September 2014

I2xz (kg m2 ) −5.776E+004 Cr (N s/m) 9080.000


I2yz (kg m2 ) 0 Ct (N s/m) 9080.000
l1 (m) 0.980 r1 (m) 0.500
l2 (m) 0.980 r2 (m) 0.500
l3 (m) 3.900 r3 (m) 0.500
D1 (m) 1.500 r4 (m) 0.500
D2 (m) 1.500 r5 (m) 0.500
D3 (m) 1.500 r6 (m) 0.500
z0 (m) −0.530 Iw1 (kg m2 ) 46.000
rpx (m) −0.300 Iw2 (kg m2 ) 46.000
rpy (m) 0.000 Iw3 (kg m2 ) 50.000
rpz (m) −0.400 Iw4 (kg m2 ) 50.000
pG1x (m) 0.000 Iw5 (kg m2 ) 50.000
pG1y (m) 0.000 Iw6 (kg m2 ) 50.000
pG1z (m) −0.3
ptx (m) −1.5
pty (m) 0.000
ptz (m) 0.000
tG2x (m) −1.739

Figure 8 shows position and velocity for different braking moments. As it can be inferred
from the graphs, when the braking moments increases, the amount of time needed to stop the
vehicle decreases. A sharp decrease in the velocity profile is observed when a huge braking
moment is applied to the system.

Figure 8. Vehicle responses to different braking moments (left, velocity; right, position).
882 M. Biglarbegian and J. W. Zu
Downloaded by [University of Sydney] at 11:56 05 September 2014

Figure 9. Velocity profiles for different braking moments for two-body model with and without fluid.

In addition, the braking moments applied to the tractor–semitrailer model were increased
in order to observe the effect of a larger amount of braking moment. When braking moments
were increased, braking time decreased. However, it turned out that in the case of tractor–
semitrailers, increasing braking moment to more than 7000 N m produced no significant
results. This is because the longitudinal forces applied from the ground to the tires can only
support a limited amount of force. This limitation is due to several factors, including the elas-
ticity of the tire, the friction between the tire and the ground, and lastly, the temperature and
pressure of the tire. Therefore, the drivers of such vehicles require a considerable amount of
time to reach a standstill from the moment they apply the brakes. Thus, the handling of such
vehicles during braking should not be underestimated.

Figure 10. Vehicle responses to different fluid levels inside the tanker (left, velocity; right, position).
Tractor–semitrailer model for vehicles 883
Downloaded by [University of Sydney] at 11:56 05 September 2014

Figure 11. Fluid forces (left, x direction; right, y direction).

Also, the dynamic behaviors of these models were studied in the absence of sloshing. It was
determined that vehicles carrying fluids need a significantly greater amount of applied braking
moment in comparison to vehicles carrying solids in order to come to a complete standstill.
This fact is depicted in figure 9 where the results are compared for the braking moments of
3000 and 7000 N m.
Subsequently, the effect of varying amounts of fluid volumes in the tanker were studied
using tanker volumes of one-half, one-third, and two-thirds. Results are shown in figure 10.
Results revealed that as the level of fluid inside the tanker increased, the sloshing forces to
the tanker’s walls increased or decreased accordingly. Thus, different dynamic behaviors were
obtained for each of these levels, which means that increased levels of fluid required more
braking moment to be applied in order to stop the vehicle. Overall, it was determined that
tankers should be specially designed to address the specific characteristics and VOF inside the
tanker.
Figure 11 shows the fluid force in x and y directions. As seen in this figure, at the moment
when the vehicle is stopped, liquid inside the tanker is still moving and exerts the sloshing
forces against the walls of the containers, but gradually the magnitude of the force decreases
until it reaches zero.

6. Conclusions

In this article, a complete dynamic model, namely tractor–semitrailer, to study the dynamic
behavior of vehicles subject to liquid slosh loads during braking and is presented. The model
includes a suspension system along with wheel dynamics. Vehicles carrying fluids feature two
subsystems: a vehicle body system and a fluid system. The two subsystems are coupled and thus
required the equations of motion to be solved simultaneously. To simplify the complicated
solution procedure, the coupled rigid and fluid equations are solved separately using two
different codes. Different braking moments are applied to the aforementioned model and
results are shown and discussed. In addition, the braking moments applied to the model
are increased in order to observe the effect of a larger amount of braking moment. When
braking moments are increased, braking time decreased. However, it turned out that in the
case of tractor–semitrailers, increasing braking moment to more than 7000 N m produced no
significant results. This is because the longitudinal forces applied from the ground to the tires
can only support a limited amount of force. Therefore, the drivers of such vehicles require
884 M. Biglarbegian and J. W. Zu

a considerable amount of time to reach a standstill from the moment they apply the brakes.
Thus, the handling of such vehicles during braking should not be underestimated. Moreover,
the dynamic behavior of the model was studied in the absence of sloshing. It was determined
that vehicles carrying fluids need a significantly greater amount of applied braking moment
in comparison to vehicles carrying solids in order to come to a complete standstill. Finally,
the effect of varying amounts of fluid volumes was studied using different volumes from one-
half to two-third of the tanker and was shown that tankers should be specially designed to
address the specific characteristics and VOF inside the tanker. It is concluded that the strategy
used in this article for solving solid–fluid interaction by incorporating vehicle dynamic effects
represents an effective method for determining the dynamic behavior of vehicles carrying
fluids in other critical maneuvers such as cornering.
Downloaded by [University of Sydney] at 11:56 05 September 2014

Acknowledgements

The authors would like to express their appreciation to ‘Simulent’ for developing the fluid
model. Financial support from the National Sciences and Engineering Research Council of
Canada is also appreciated.

References
[1] McCarty, J.L. and Stephens, D.G., 1960, Investigation of the natural frequencies of fluids in spherical and
cylindrical tanks. NASA TN D-252.
[2] Ranganathan, R., Ying, Y. and Miles, J.B., 1994, Development of a mechanical analogy model to predict the
dynamic behavior of liquids in partially filled tank vehicles. Society of Automotive Engineers (SAE), 942307.
[3] Rakheja, S. and Wang, Z., 1996, Analysis of braking process of a partially-filled tractor tank semitrailer.
Advances in Transportation Systems (CSME Forum SCGM), pp. 326–333.
[4] Ranganathan, R. and Ying, Y.S., 1996, Impact of liquid load shift on the braking characteristics of partially filled
tank vehicles. Vehicle System Dynamics, 26, 223–240.
[5] Ranganathan, R., Rackheja, S. and Sankar, S., 1989, Steady turning stability of partially filled tank vehicles with
arbitrary tank geometry. Transactions of the ASME, Journal of Dynamic Systems, Meausurment and Control,
111, 481–489.
[6] Rakheja, S., Sankar, S. and Ranganathan, R., 1988, Roll plane analysis of articulated tank vehicles during steady
turning. Vehicle System Dynamics, 17, 81–104.
[7] Ranganathan, R., Rakheja, R. and Sankar S., 1990, Influence of liquid load shift on the dynamic response of
articulated tank vehicles. International Journal of Vehicle Mechanics and Mobility, Vehicle Systems Dynamics,
19(4), 177–200.
[8] Rakheja, S. and Ranganathan, R., 1993, Estimation of the rollover threshold of heavy vehicles carrying liquid
cargo. Heavy Vehicle Systems International Journal of Vehicle Design, 1, 79–98.
[9] Papov, G., Sankar, S. and Sankar, T.S., 1996, Shape optimization of elliptical road containers due to liquid load
in steady-state turning. Vehicle System Dynamics, 25, 203–221.
[10] Acarman, T. and Ozguner, U., 2003, Rollover prevention for heavy trucks using frequency, shaped sliding mode
control. IEEE, 1, 7–12.
[11] Sankar, S., Ranganathan, R. and Rakheja, S., 1992, Impact of dynamic fluid slosh loads on the directional
response of tank vehicles. Vehicle System Dynamics, 21, 385–404.
[12] Rumold, W., 2001, Modeling and simulation of vehicles carrying liquid cargo. Journal of Multibody System
Dynamics, 5, 351–374.
[13] Parizi, H., Fard, M.P. and Dolatabadi, A., 2003, A computer code to design liquid containers for vehicles. CFD
conference, Vancouver.
[14] Gillespie, T.D., 1992, Fundamentals of Vehicle Dynamics (SAE).
[15] Wong, J.Y., 2001, Theory of Ground Vehicles (3rd edn) (John Wiley & Sons Inc).
[16] Bakker, E., Pacejka, H.B. and Lidner, L., 1989, A new tire model with an application in vehicle dynamic studies.
Society of Automotive Engineers (SAE), paper 890087.
[17] Pacejka, H.B., 2002, Tyre and Vehicle Dynamics (1st edn) (Butterworth Heinemann).
[18] Ginsberg, G., 2000, Advanced Dynamics (2nd edn) (John Wiley & Sons, Inc.).
[19] Dogoff, H. and Murphy, R.W., 1971, The dynamic performance of articulated highway vehicles – a review of
the state-of-art. SAE Transaction. No. 710223.
[20] Mikulcik, E.C., 1971, The dynamic of tractor semitrailer vehicles: the jackknifing problem. SAE paper
No. 710045.
[21] Bernard, J., 1974, A digital computer method for the prediction of the directional response of trucks and
tractor-trailers. SAE Paper No. 740138.
Tractor–semitrailer model for vehicles 885

[22] Vlk., F., 1982, Lateral dynamic of commercial vehicles. Vehicle System Dynamics, 11, 305–324.
[23] Ellis, J.R., 1983, A model of semitrailer vehicle including roll modes. Proceedings of the IUTAM Symposium
on Roads and on Tracks, Vienna, pp. 184–202.
[24] Ervin, R.D., 1983, The influence of size and weight variables on the roll stability of heavy duty trucks. SAE
Paper, No. 831163.
[25] El-Gindy, M. and Wong, J.Y., 1987, A comparison of various computer simulation models for predicting the
directional response of articulated vehicles. Vehicle System Dynamics, 16, 249–268.
[26] Fancher, P.S. and Bareket, Z. 1992, Behavior of articulated vehicles on curves, heavy vehicles and roads:
technology, safe, and policy. Proceedings of the Third International Symposium on Heavy Vehicles and Weights
and Dimensions, Cambridge, pp. 331–337.
[27] Chen, C., 1996, Backstepping design of nonlinear control systems and its applications to vehicle control in
automated highway system. PhD dissertation, University of Berkeley.
[28] Chen, C. and Tomizuka, M., 1995, Dynamic modeling of articulated vehicles for automated highway system.
Proceedings of American Control Conference, Seattle.
[29] Chen, C. and Tomizuka, M., 1995, Dynamic modeling of tractor–semitrailer vehicles in automated highway
system. California PATH Research Report, UCB-I TS-PWP-95-8, Berkeley.
Downloaded by [University of Sydney] at 11:56 05 September 2014

[30] Hyun, D., 2001, Predictive modeling and active control of rollover in heavy vehicles, PhD dissertation, University
of Texas A&M.
[31] Bussman, M., Mostaghimi, J. and Chandra, S., 1999, On a three-dimensional volume tracking model of droplet
impact. Physics of Fluids, 11(6).

You might also like