Vous êtes sur la page 1sur 21

Finite Elements in Analysis and Design 31 (1999) 295—315

Vehicle dynamics simulations with coupled multibody


and finite element models
C.W. Mousseau *  , T.A. Laursen, M. Lidberg, R.L. Taylor
Transportation Research Institute, University of Michigan, 2901 Baxter Road, Ann Arbor, MI 48109, USA
 Department of Civil and Environmental Engineering, Duke University, USA
 Mechanical Dynamics Inc., USA
 Department of Civil and Environmental Engineering, University of California, Berkeley, USA

Abstract

With the advent of multibody system simulations (MSS) programs, it has become common practice to use computer
modeling to evaluate vehicle dynamics performance. This approach has proved to be very effective for predicting the
handling performance of vehicles; however, it has proved less successful for predicting the vehicle response at frequencies
that are of interest in ride harshness and durability applications. The lack of correlation between theory and experiment
can be partially traced back to tire models that are inadequate for rough road simulation. This paper presents
a comprehensive vehicle dynamics model for simulating the dynamic response of ground vehicles on rough surfaces. This
approach uses a MSS program to simulate the vehicle and a nonlinear FE program for the tires. Parallel processing of the
tire models improves the efficiency of the overall simulation. Applications for this technology include vehicle ride and
harshness analysis and durability loads simulation. This paper describes the MSS vehicle model, the tire FE model, and
the interface which transfers data between the two simulations. Simulation and experiment results for a single tire
without a vehicle encountering an obstacle and for a vehicle with four tires driving across a pot hole are presented.
Conclusions and opportunities for further research end the paper.  1999 Elsevier Science B.V. All rights reserved.

Keywords: Vehicle dynamics; Tire mechanics; Durability; Friction; Surface contact

1. Introduction

With the advent of multibody system simulations (MSS) programs, it has become common
practice to use computer modeling to evaluate vehicle dynamics performance. This approach has
proved to be very effective for predicting the handling performance of vehicles; however, it has
proved less successful for predicting the vehicle response at frequencies that are of interest in ride

* Corresponding author.
 Formerly with the Ford Motor Co.

0168-874X/99/$ — see front matter  1999 Elsevier Science B.V. All rights reserved
PII: S 0 1 6 8 - 8 7 4 X ( 9 8 ) 0 0 0 7 0 - 5
296 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

harshness and durability applications. For instance, a rigid-body vehicle model may prove
adequate for predicting the handling response on flat surfaces. However, it may prove inadequate
for simulating a vehicle traveling over irregular terrain, since high-frequency modes may be excited
in the vehicle and tire structure. Also, the vast majority of vehicle handling simulations are
conducted on flat surfaces, and the tire is supported at the ground by only a single point. In reality,
contact between a tire and ground occurs over a distributed area, requiring a more sophisticated
modeling approach.
Because of the continuous nature of the deformation and contact, experience has shown that
tires are best modeled using the nonlinear finite element (FE) method. MSS programs tends work
well for simulating systems of interconnected rigid elements, but become problematic when contact
and large structural deformations occur. On the other hand, it is more difficult to build rigid-body
models with nonlinear FE programs than with MSS codes. Libraries of joint constraints, discrete
force elements, and built in subroutines make the MSS approach more straightforward to use. One
solution to this dilemma is to incorporate both types of methodologies into a single simulation. In
this approach we simulate the vehicle with MSS program and tires with nonlinear FE programs.
This approach was demonstrated with a simple two-degree-of-freedom (DOF) vehicle dynamics
model and a single-tire finite element (FE) model [1]. However, the vehicle simulated in that study
was very generic; therefore, the results could not be directly applied to a production engineering
application. Also no experimental data were available to validate the combined vehicle and tire
simulation.
This paper presents a comprehensive vehicle dynamics model for simulating the dynamic
response of ground vehicles on rough surfaces. This approach uses a MSS program to simulate the
vehicle and a nonlinear FE program for the tires. Parallel processing of the tire models improves
the efficiency of the overall simulation. The paper is organized as follows. The multibody vehicle
dynamics model, tire model, and interface, which controls the program execution and the flow of
data, are described. Simulation and experiment results for a single tire without a vehicle encounter-
ing an obstacle and for a vehicle with four tires driving across a pot hole are presented. Conclusions
and suggestions for future work close the paper.

2. Simulation models

2.1. Vehicle dynamics model

A MSS model of a mid-size automobile with a unitized body, a MacPherson strut [2] front
suspension, and a multi-link [2] rear suspension was constructed for this study. Fig. 1 illustrates
the front and rear suspensions. This figure shows that the suspension components (e.g., control
arms and spindle) are connected together with a combination of ball joints (i.e., rigid metal to metal
connections), and bushings (i.e., relatively complaint elastomeric elements). Bushings and a sub-
frame help to isolate the body from vibrations that originate in the front suspension. The engine,
which is connected to the body via elastomeric mounts, is also an important contributor to the
dynamic response of the vehicle.
The steps involved in constructing the MSS model include: (1) identifying the vehicle compo-
nents that have significant mass (e.g., car body, spindle, and control arms), (2) identifying how the
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 297

Fig. 1. (a) Front and (b) rear vehicle suspensions.

components connect together (i.e., system topology), and (3) determining the nature of forces
exerted on the components. We assume that components with significant masses can be adequately
represented by rigid bodies [3,4]. For instance, the vehicle body, sub frame, engine, controls arms,
and spindle are modeled as rigid bodies. Note this approach works well for components that are
298 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

fairly rigid (e.g., control arms), and consequently, have natural frequencies that are far above the
excitation frequencies. However, errors may result if significant structural deformation occurs or
critical modes of vibration become excited [5].
Fig. 2 illustrates the left front suspension portion of the MSS vehicle dynamics model. Each
element in the figure represents a particular dynamic process and corresponds directly to a suspen-
sion element. This figure shows that constraint equations, represented pictorially by joints [3,4],
are used to model rigid connections between components. For example, a ball joint is modeled by
imposing a constraint requiring two points on two bodies to be coincident during the analysis.
Likewise, bushing elements [3] are used to model compliant connections between components.
These elements use forces that are solely function of displacement and velocity to link bodies, to
maintain the connection, allowing for deformations to occur during the analysis. Shock absorbers,
springs, and jounce stops are represented with forces that depend nonlinearly on the relative
displacement and velocity between two points [3,4].
The MSS’s described in this paper were generated with the ADAMS [3,4] program. The data file
that describes the simulation model, is constructed from the system topology; mass and inertia
properties of vehicle body and suspension components; the force-deflection characteristics of the
bushings and springs; the force-velocity characteristics of shock absorbers. The complete MSS
vehicle model has 32 rigid bodies, 32 joints, 26 bushing elements and 17 spring-damper elements.
ADAMS uses this information to numerically assemble and solve the dynamic equations of motion
of the system.

Fig. 2. Vehicle dynamics model left front suspension topology.


C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 299

2.2. Tire model

The deformation of the tire structure and contact between the tire and ground is modeled using
the FE method. The tire FE described below is implemented as user element in FEAP [6,7],
a research FE program. Since large deformations occur in the tire structure and discontinuities are
present in contact model, the resulting equations of motion are nonlinear. Thus the quasi-static and
dynamic solutions are generated by the Newton Raphson [6], Newmark [6], and HHT [6]
nonlinear, implicit solution procedures in FEAP.
A brief description of the tire FE model is provided below. The tire structure model, illustrated in
Fig. 3, consists of 50 to 100 wedge-shaped finite elements equally spaced along the circumference of

Fig. 3. Tire finite element model.


300 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

the tire. The conductivity of each element is defined by three nodes. The center node defines the
spindle coordinates and the other two describe tread coordinates. Each node has 3 degrees of
freedom (DOF) corresponding to vertical displacement along the ½-axis, horizontal displacement
along the X-axis, and rotation about the Z-axis. The tire rim rotation is described by the rotation
of the spindle node. A geometrically nonlinear beam element that accommodates large deforma-
tion [8] models the tread, while the circular arch model approximates the sidewall behavior.
Membrane forces from the arch provide both radial and circumferential support continuously
along the tread. The circular arch introduces the effect of sidewall bulging into the model to
improve accuracy under large deformations.
Fig. 4, shows a cross section of the tire element during deformation and demonstrates it
operation. Initially, the element is unloaded, and the tread and sidewall are at their relaxed
configurations. The equilibrium configuration after pressurization depends on these stiffnesses and
the tread width, and the sidewall radius. Application of a vertical load on the tread nodes causes it
to move vertically, and the sidewall radius to decrease until a new equilibrium point is reached. The

Fig. 4. Deformation of tire finite element.


C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 301

figure demonstrates that even though the tire finite element is two dimensional, it does incorporate
the effect of sidewall bulging.

2.2.1. Sidewall deformation


The sidewall provides support for the tread in both the circumferential and radial directions.
One way to visualize how the sidewall works, is to assume that the tread is linked to the rim by
a bundle of inextensible fibers, a concept very similar to the cords of a radial ply tire. Circumferen-
tial forces are added to model to account for the shearing resistance of the sidewall rubber-cord
composite. Since the model is 2D, the two inextensible circular membranes act in tandem, so no
lateral motion is allowed. For a given cross section, the kinematics of the two sidewall cords are
defined by an upper point which is attached to the rim, and the lower point which is attached to the
tread. The upper point displacement is determined by the displacement and rotation of the rim
node. The lower point displacement is determined by the displacement of the two nodes that define
the tread.
We make the following assumptions about the behavior of the sidewall: (1) the elastic character-
istics of the arch are assumed to be independent of h; (2) the membrane analogy still applies; (3) the
additional force that arises from elastic deformation of the sidewall can be approximated using
a circular arch (i.e., beam) model; and (4) the direction of the additional force is determined by the
membrane model. With these assumptions, the forces acting at the sidewall/tread interface contains
two components: a membrane force and a force resulting from the elastic deformation of the arch.
The sidewall force relationship that results is both sufficiently accurate and easy to compute.
Since bending alters the nature of the deformation field, the arch will no longer maintain a true
circular shape. However, the sidewall bending stiffness is relatively low and membrane effects
dominate. The resulting shape is very close to a circular arc and the arch radius is independent of
the meridian angle. Given that the arch remains circular, from Fig. 5, the meridian angle h can be

Fig. 5. Forces acting on ideal sidewall geometry.


302 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

determined from the length of the sidewall (l ) and the sidewall section height (¸) as


 
l ¸ p
 cos h# h! "0. (1)
2 2 2
This relation is needed for computing the sidewall’s contribution to the tangent stiffness matrix.
Also of use is the relation between the deformed radius r, h and ¸:
¸
r" . (2)
2 cos h
The following expression for vertical projection of the sidewall support force acting on the tread
was developed in [15],

  
1 1
¹" ! k #rp sin h. (3)
 r rr P R

Eq. (3) depends on the empirically derived effective sidewall bending stiffness (k ) and the unde-
P
formed sidewall radius (r ).

The model as it stands, relies on horizontal projection of the membrane force to provide
resistance to shear. The additional resistance to shear that sidewall rubber-cord composite
provides needs to be accounted for in the model. We use an empirical model to approximate the
additional shear forces that act on the rim and the tread. We assume that a force to be proportional
the shear angle c, and it acts on the tread and rim in circumferential direction in an equal and
opposite manner.
F "k c, (4)
V A
where, k is the effective sidewall shear stiffness.
A
2.2.2. Element kinematics
To derive the FE tangent stiffness matrix and force residual, we need to determine the force
acting on an arbitrary cord. The force per unit length acting on the tread and rim is determined by
the cord length, cord orientation, and inflation pressure. Fig. 6 shows a view in the X—½ plane of
the element in undeformed position and in the deformed position, (i.e., with the rim rotated by
h and with the tread nodes deformed). The direction of the distributed force that acts between the
rim and the tread is assumed to lie along the direction of the cord (e ). Since the sidewall must
W
maintain a circular shape, the membrane radius will change as the two points move closer together,
causing the force acting on the tread to also change. This simulates the effect of sidewall bulging.
To determine the cord force, we need to determine the length and direction of a sidewall cord in
terms of the displacements (u , u , and u ) and reference coordinates (X, ½). Referring to Fig. 6, the
V W F
vector that defines the fiber in the deformed condition is given by
P "P #P !P "(X #u )i#(½ #u ) j#r sin(
!
#h)i
YY Y YY RY  V  W  
!r sin(
!
#h) j!(X #u )i!(½ #u ) j. (5)
   V  W
The horizontal projection of the sidewall cord is given by
¸"""P """(P · P (6)
YY YY YY
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 303

Fig. 6. Deformation of a single tire finite element.

The unit vector e is aligned along the deformed sidewall cord.


W
P
e " YY . (7)
W ""P ""
YY
The unit vector e lies in the X—½ plane and is normal to the deformed sidewall fiber.
V
e "k;e . (8)
V W
We are also interested in determining the forces that the cord act on the tread and react at the
spindle point. Recall that the cord illustrated in Fig. 6 is actually a projection of the circular arch in
the X, ½ plane. The force acting between the sidewall and tread consists of a membrane component
and an elastic component due bending that acts along the e direction, and a shear component that
W
acts along the e direction. Eqs. (3) and (4) are used to express the force in vector form.
V

  
1 1
T " p r#k ! sin he #k ce . (9)
  P r rr W A V

Referring again to Fig. 6, the forces and moment that act at the spindle point S are given by

  
1 1
T "! p r#k ! sin he #k ce , (10)
  P r rr W A V


  
1 1
M "! p r#k ! sin hp ;e !jk p ;e . (11)
  P r rr YY W A YY V

Note that we are assuming that the sidewall fiber cannot support a moment about the Z-axis.
304 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

2.2.3. Sidewall element internal forces and tangent matrix


Our intention is to use implicit methods to solve for the displacements during both the
quasi-static and dynamic analysis. This requires the formation of the element tangent stiffness
matrix and force residual. The tangent stiffness matrices for the implicit solution procedures are
found by constructing and linearizing the weak form [6,7]. At this point it is convenient to use
Eqs. (9)—(11) to form the following residual vector
u(u)"+T · i T · j 0 T · i T · j M · k, (12)
    
The weak form for the distributed force is formed by integrating over the element interval, the
product of the residual and a vector of weighting functions (w) [6].


F
G (u)" u(u) · w dX. (13)


The weak form is linearized about a displacement u and the finite element matrices are formed with
the Galerkin method [7]. In this step, the tread, spindle nodal point positions, and nodal
displacements are interpolated isoparametricly [7]. The element tangent stiffness matrix is found
by substituting in the interpolated variables and the stiffness matrix is obtained by factoring out the
weighting coefficients and nodal displacements. This process is described in more detail in [9].

2.2.4. Material damping


The primary source of energy dissipation in a rolling tire is hysteresis in the tread and the
sidewall. However, hysteresis can be difficult to implement and significantly increase the amount of
computation required in a finite element simulation. Therefore, the energy dissipation is calculated
in the model with equivalent viscous damping. The damping coefficient is obtained by equating the
energy dissipated during a sinusoidal deformation of a material specimen to the energy dissipated
by a viscous damper, as described in [10]. There are three sources of damping in the simulation
model, all of which use a linear damping law. In the tread, the damping moment is assumed to be
proportional to the rate of change of the curvature. We assume that the sidewall damping acts in
the radial and circumferential directions.

2.2.5. Contact model


The mathematical constraints involved in defining the road/tire mechanical interaction can be
generally understood by referring to Fig. 7. In the figure, X refers to the reference position of a node
on the periphery of the tire, and the contact kinematic quantities associated with X serve to
quantify its motion relative to the opposing roadway surface. The first of these is the gap function g,
which can be written as:
g(X)"([X#u(X)]![YM #u(YM )]) ) l(YM ), (14)
where YM is the closest point projection of X on the surface and l is the surface normal at this location.
Impenetrability of tire and roadway is then mathematically expressed as
g(X))0 (15)
for all nodes X on the periphery of the tire. It is further assumed that all mechanical interaction
normal to the interface is compressive, so that no adhesive forces form between the tire and
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 305

Fig. 7. Penetration calculation for contact model.

roadway. In view of this fact, the relationship between the contact pressure t (assumed positive in
,
compression) and the gap function g is expressed via the following classical Kuhn—Tucker
conditions:
g(X))0, t (X)*0, t (X)g(X)"0. (16)
, ,
Expressed in this form, we can think of t as being the Lagrange multiplier needed to enforce
,
inequality condition (15) for node X.
In the model we propose, frictional interaction is described using a Coulomb law. For a node
undergoing persistent sliding, where gR "0, a frame indifferent measure of relative sliding is given by
a projection of the relative material velocity into the plane tangent to the roadway surface
V (X)"(I!ll) (V(X)!V(YM )). (17)
2
A Coulomb description governing the tangential stress t is given in terms of a slip function U via
2
t
: ""t ""!kt )0, V (X)"1 2 , 1*0, 1U"0.
U(t , t ) " (18)
2 , 2 , 2 ""t ""
2
Eq. (18) imply a law of constrained evolution for the frictional stress t , in a manner largely
2
analogous to the way rate independent plasticity equations are often formulated. The familiar
physical assumptions of Coulomb are readily confirmed: the magnitude of frictional stress ""t "" may
2
not exceed the coefficient of friction k times the pressure t , all sliding occurs in a direction
,
opposing the applied frictional stress, and no sliding occurs unless f'0, which is only allowed if
""t "" is exactly equal to kt .
2 ,
With respect to the global finite element equations, contact effects appear as
Md$ (t)#F (d)#F"0, (19)
where the first term represents the dynamic effects, F represents the internal forces generated by
tire deformation as discussed above, and F is the contact force vector, assembled from the contact
tractions satisfying conditions (16) and (18). Of course, F is an extremely nonlinear and discontinu-
ous function of the tire deformation. In particular, contact conditions are widely recognized to be
306 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

subject to ill-conditioning if penalty methods are used. Accordingly, we use the augmented
Lagrangian method as documented in [11] to enforce the aforementioned constraints.

2.2.6. Tire model parameters


The model parameters are organized into a group related to tire geometry and another
associated with elastic properties. The parameters that describe the tire geometry, including, the
uninflated tire diameter, effective rim diameter, tread width, and sidewall length are measured
directly from the tire cross section. Other parameters such as inflation pressure and surface friction
are specified by the user. Since the circular membrane model only approximates the actual tire
cross-sectional shape, the user must exercise some judgment when interpreting the sidewall length
and tread width parameters. The axial stiffness is best measured directly with a uniaxial test. The
bending stiffness can be estimated either with a three point bending test on a specimen cut from the
tire, or from load deflection tests on a flat surface and knife edge with the complete tire.
To enable the model to better simulate the response of the tire during large deformations, an
empirical term that approximates bending in the sidewall is added. Since the elastic sidewall
bending stiffness (k ) parameter is difficult to measure directly, an indirect approach was used. For

our studies, it was determined by matching simulations to force—deflection experimental data at
a single inflation pressure. These simulations took less than a minute to run, and a suitable value
was determined after only five simulations. The parameter is independent of inflation pressure, and
thus data are required only for a single inflation pressure to characterize the model.

2.3. Interfacing the nonlinear finite element multibody dynamics simulations

To conduct a full vehicle simulation, the MSS vehicle dynamics model requires forces for each of
the four spindle locations and each FE tire model requires the corresponding spindle displace-
ments. These data needs to be passed between the simulation during every simulation time step.
The interface needs to perform this task both reliably and efficiently.

2.3.1. MSS simulation data requirements


Using information from a data file, the ADAMS program assembles and solves the Eu-
ler—Lagrange dynamic equations in first order form. A predictor—corrector algorithm with step size
control integrates the resulting set of differential algebraic equations (DAEs). A backward differ-
ence formula (BDF) approximates the solution over an interval and transforms the DAEs into a set
of coupled implicit algebraic equations [12,13]. The following BDF formula is used:
I
q (t )"hb qR (t )# (a q (t )#hb qR (t )). (20)
G J  G J\H H G J\H H G J\H
H
The ith generalized coordinate (q) at time step l is estimated from a linear combination of the
generalized coordinate at the previous kth time steps and its derivative. Where, a and b are
weighting coefficients and h is the time step size. The predictor provides an initial estimate of the
solution, while the corrector uses the BDF and a Newton-Raphson like algorithm to iteratively
solve for the final value of q (t ) [13]. At each iteration of the solver, this procedure requires
G J
evaluation of the forces and formation of the Jacobian matrix (i.e., the partial derivative of the
forces with respect to the generalized coordinates).
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 307

2.3.2. Interface operation


Fig. 8 illustrates the major computations that take place during the simulation and the data flow
between processes. During each iteration of the MSS corrector, the vehicle dynamics model
provides the spindle position to tire model simulation which implicitly solves the nonlinear tire FE
problem for the spindle force. Since both simulations are using implicit methods to solve their
respective problems, many iterations of nonlinear FE solver are required to generate a simulation.
For example, a vehicle dynamics simulation may last 10 or more second, and the each time step is
about 1 ms, the simulation will require about 100 000 iterations of the FE solver.
Data transfer between processes is accomplished with library calls to the transport layer
interface (TLI) [14]. This software allows for data to be transferred reliably between processes that
may exist on a single computer, or on a group of networked computers. At the start and end of each
iteration of the MSS corrector, only spindle forces and displacements pass back and forth between
the MSS and the FE programs. When compared to the simulation as a whole, this is a very small
amount of data. With the exception of this data, the calculations in a given tire model are
independent of the other simulations. Thus, two or more tire simulations can execute concurrently,
thereby, further reducing the simulation run time.

Fig. 8. Simulation model interface.


308 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

2.3.3. Force Jacobian calculation


The convergence of the ADAMS solver can accelerated if, in addition to the spindle force, the
Jacobian matrix is provided. The Jacobian matrix can be calculated either directly from the FE
program, or by numerically perturbing the spindle position. The former approach generates the
Jacobian matrix as a by product of the tangent stiffness matrix calculation, the latter requires
additional evaluations of FE program. Since both evaluating the tire finite element model and
moving data between simulations is very time consuming, and can lead to longer simulation run
times, the direct method is preferred. The other benefit of the direct method is that the resulting
matrix is consistent [6]. This property increases the rate of convergence; therefore, further
improves the efficiently of the simulation.
The procedure we used to extract the spindle force Jacobian from FE tangent matrix is described
below. For every iteration of the ADAMS corrector we need to calculate the spindle force, which
depends on the acceleration of tread mass and damping of the tire structure. The Jacobian matrix
found by differentiating the spindle force with respect to the relevant spindle displacements:
*F (u, uR , u( )
F " G . (21)
G H *u
H
Note that we are dealing with a force that is a function of displacement, velocity, and acceleration
of all the nodes in the tire structure. Using the chain rule, we further carry out the differentiation:
*F *F *uR *F *u(
F " G# G H# G H . (22)
G H *u *uR *u *u( *u
H H H H H
Using Newmark predictors [6] to extrapolate the velocity and acceleration in the above equation,
the following expression results:

   
*F c *F 1 *F
F " G# G# G. (23)
G H *u hb *uR hb *u(
H H H
Defining the following quantities:
*F *F *F
K " G, C " G and M " G ,
GH *u GH *uR GH *u(
H H H
we can restate Eq. (23) as

   
c 1
F "K # C # M , (24)
G H GH hb GH hb GH

where h is the time step size and b and c are the Newmark parameters. In Eq. (24) K , C , and M ,
GH GH GH
are loosely described, respectively, as the “effective” stiffness, damping, and mass matrices. Eq. (24)
also shows that the force Jacobian (F ) depends on the duration over which the perturbation
G H
occurs. The matrix on the right-hand side of Eq. (24) is identical in form to the stiffness matrix in
the Newmark integration algorithm [6].
To calculate the spindle force Jacobian we sub-divide the tire model FE tangent matrix and force
residual into parts containing the spindle node (i.e., the subscript s) and not containing the spindle
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 309

node (i.e., the subscript u).

    
K K Du R
   "  . (25)
K K Du R
  Q 
The tangent matrix can be partially factored into

     
K K L 0 D 0 U U
  "     . (26)
K K L I 0 H 0 I
   
The matrix H , the reduced tangent matrix for the spindle node, is found by substitution to be:

H "K !L D U . (27)
    
The computation of the spindle force that is exported to the MSS program is shown below:
f "R !L L\R , (28)
    
where the inverse is done as a forward solution of following the equations.
L w "R (29)
  
Thus,
f "R !L w . (30)
   
Eqs. (27)—(30) were implemented into the FEAP program as a macro command which is called
when the spindle force is required by the vehicle MSS.

3. Results

Simulation results from both the standalone tire model and the combined vehicle and tire models
are presented below. Experimental results are also presented to demonstrate the accuracy of the
simulation models.

3.1. Tire model alone

To demonstrate the accuracy of the tire model for predicting impact forces, simulation results are
compared to experimental data. The tire FE simulations were conducted with the FEAP program
using the Newmark dynamic solution procedure. The FE model consists 100 tire FE s equally
spaced along the tire circumference. Contact between the tire and ground was modeled with 2D
surface contact elements. Fig. 9 illustrates the simulation model at the start of the dynamic
analysis; in the inflated condition loaded against a simulated rigid drum. The rigid contact surface
has a 1700 mm diameter with a step that is 25 mm high by 660 mm long, and the “x” in the center of
drum defines the position of a node to which the rigid surface is constrained. A coefficient of friction
of 0.3 was defined for the surface.
The total duration of the simulation was 8 s and it consisted of two parts. The first part was
quasi-static and it lasted 2 s; the latter part was dynamic. During the first part, the tire was inflated
310 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

Fig. 9. Tire model at start of simulation.

to 0.208 MPa and deformed with a load of 3800 N against the surface. During the second part of
the simulation, the drum was spun from rest to the final angular velocity over 4 s. The combination
of the surface movement and friction caused the tire to start rolling. After 4 seconds the drum
velocity was then held constant and the tire was allowed to reach a steady-state condition before it
impacted the obstacle. The simulation execution time was 1.3 CPU hours on an HP 735
workstation.
The experiment was conducted on a 1700 mm diameter road wheel with a 25 mm high by
660 mm long obstacle clamped to the outside surface. The tire was mounted to a dead axle which
was instrumented to measure longitudinal and vertical forces. The drum was spun up to speed with
tire rolling against the road wheel surface. A data acquisition system captured the spindle force
response when the tire impacted the obstacle. More details of the tire impact test are described
in [15].
Fig. 10 shows the measured and predicted hub forces for the tire rolling over the obstacle at
51.3 KPH. The measured and simulated longitudinal force responses agree reasonably well.
However, the model underestimates the peak values slightly and predicts a slightly higher
frequency of oscillation at 44 Hz, compared to 35 Hz observed in the test. The 44 Hz oscillation
corresponds to the first natural frequency of the tire model when loaded against a rigid surface
[15]. The reason for differences between the model and test frequencies are described in [15] and
stem largely from the simplifying assumption used to model the tread. Very similar results were
obtained with the Newmark and HHT integration algorithms.
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 311

Fig. 10. Simulated and measured (a) longitudinal and (b) vertical spindle forces for the 145 SR-12 tire impacting a step on
a 25 mm high step rotating drum at 51.3 KPH (31.9 MPH).

The vertical spindle force response shows very good agreement between the experimental results
and simulation. The simulated response shows a 89 Hz oscillation when the tire reaches the top of
the step, and an 84 Hz oscillation when the tire leaves the step. These oscillations are attributed to
the first vertical mode of the tire that varies slightly with tire deflection [15]. The measured
response shows a 88 Hz oscillation when the tire reaches the top of the step, and an 85 Hz
oscillation when the tire leaves the step. This exercise was repeated at different speeds and the peak
values predicted by the simulation show good agreement with the experimental results [1].

3.2. Combined tire and vehicle dynamics models

To explore accuracy of the combined tire and vehicle simulation, a limited validation study was
carried out. An experiment was conducted on a vehicle instrumented with wheel force transducers
312 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

that measured the longitudinal, lateral, and vertical spindle forces. The experiment was conducted
in the following manner. The vehicle was initially driven over a flat surface at a constant speed of
48 KPH where it then encountered a 760 mm long by 100 mm deep obstacle on the left side of the
vehicle and a flat surface on the right side. During the impact with the obstacle, time history data
from all the wheel transducers were digitized and stored for later analysis.
The analytical results were generated with the combined tire and vehicle simulation model. The
simulation was conducted with the ground moving underneath the vehicle at a prescribed velocity,
the tires rolling freely, and no tractive torques were applied to the tires. The simulation was
conducted in the following manner. During the first second, the tire was quasi-statically inflated to
0.243 MPa. During the next second, the vehicle and tires were allowed to settle out dynamically to
an equilibrium position. Over the following 5 s, the ground was accelerated at constant rate from
0 to 48 KPH. For the last two seconds of the simulation, the ground was moved at constant speed
of 48 KPH, during which the vehicle encountered the obstacle. The simulation lasted a total of 9 s
and 13.5 CPU h were required to generate the results on a two processor, shared memory SGI
workstation.
Fig. 11 shows the experimental and simulated spindle forces for left front wheel during impact
with the obstacle. This figure shows that in the vertical direction, the simulation agrees well with
the test results. Peak values are within 1% of one another and many of the oscillations exhibited by
the simulation are also present in experiment. The magnitude of the FFT of vertical force in Fig. 13
shows that the frequency of oscillation of the first vertical tire mode predicted by the simulation
agrees very well with the test results. It also shows that the vehicle dynamics model also over
predicted the frequency of hop mode at 11.5 Hz;2.7 Hz. However, Fig. 11 also shows that the
longitudinal forces predicted by the simulation over estimates the magnitude of the longitudinal
force. It also shows that the simulation under estimates the duration of the force during the initial
impact.

Fig. 11. Measured and simulation left front (a) longitudinal and (b) vertical spindle forces for an automobile traveling
48 KPH over a 0.76 m long by 0.1 deep chuckhole.
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 313

Fig. 12. Measured and simulation left rear (a) longitudinal and (b) vertical spindle forces for an automobile traveling
48 KPH over a 0.76 m long by 0.1 deep chuckhole.

Fig. 12 shows the experimental and simulated longitudinal and vertical spindle forces for left
rear wheel during the impact with the obstacle. This figure also shows good agreement between the
simulation and test in the vertical direction. However, the simulation underestimates the magni-
tude of the longitudinal force by almost 50%. Although, the duration time of the initial impact
force predicted by the simulation agrees well with the experimental results.
314 C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315

Fig. 13. FFT of the left front (a) vertical and (b) longitudinal spindle forces.

4. Conclusions

An approach for coupling a vehicle dynamics model to tire model to from a comprehensive
simulation was presented. This approach uses a MSS program to simulate the vehicle and
a nonlinear FE program for tires. The efficiency of the overall simulation is improved by using
a special purpose FE to model the tire structure and by parallel processing the nonlinear FE
programs.
The results of this study indicates that this approach is a viable technique for simulating vehicle
dynamic response on rough roads. The tire model alone agrees well with experimental data and
executes in a reasonable amount of time. Although, the combined tire and vehicle simulations
execute much more slowly than either simulation alone. direct computation of the force Jacobian
can improve the computational efficiency. For the case of the a vehicle driving over an obstacle, the
combined tire and vehicle simulation was able to predict the vertical spindle forces very accurately.
However, the longitudinal spindle forces predictions were less accurate. The sources of the
disagreement between simulation and experiment are currently being investigated.
Future work will include addressing the cause of the discrepancies in longitudinal force
predictions, increasing the efficiency of the interface between, and conducting a more comprehens-
ive validation study. Work is already underway on a 3D version of the tire model [10,16] that is
also capable of predicting lateral loads.

Acknowledgements

The authors wish to acknowledge the support of the Research Staff, Light Truck, and Car
Chassis Engineering activities of Ford Motor Co. Also we wish to acknowledge the encouragement
C.¼. Mousseau et al. / Finite Elements in Analysis and Design 31 (1999) 295—315 315

and contributions of Jim Avouris, Vikas Chawla, Sam Clark, Al Conle, Ian Darnell, Greg Hulbert,
John Hogan, Bill Oates, Steve Riley, and Natarajan Saravanan. Also, Tod Laursen was supported
in part through NSF CAREER award number CMS-9703356; this support is gratefully acknow-
ledged.

References

[1] C.W. Mousseau, M.W. Sayers, The effect of tire and suspension dynamics on wheel spindle forces, Heavy Vehicle
Systems — J. Vehicle Des., accepted.
[2] K. Newton, K. Steeds, T.K. Garrett, Motor Vehicle, Butterworths, Stoneham, MA, 1989.
[3] MDI Inc., ADAMS/Solver Reference Manual, MDI Inc., Ann Arbor, MI, 1994.
[4] N. Orlandea, M.A. Chace, D.A. Calahan, A sparsity-oriented approach to the dynamic analysis and design of
mechanical systems, parts I and II, J. Eng. Ind. 99 (1977) 773—784.
[5] W.-H. Shyu, Quasi-static mode compensation for component mode synthesis of dynamical systems, Ph.D.
Dissertation, University of Michigan, 1996.
[6] O.C. Zienkiewicz, R.L. Taylor, The Finite Element Method, vol. 2, McGraw-Hill, New York, 1995.
[7] O.C. Zienkiewicz, R.L. Taylor, The Finite Element Method, vol. 1, McGraw-Hill, New York, 1989.
[8] J.C. Simo, L. Vu-Quoc, On the dynamics of flexible beams under large overall motions — the plane case: Parts I and
II, J. Appl. Mech. 53 (1986) 849—863.
[9] C.W. Mousseau, G.M. Hulbert, An efficient tire model for the analysis of spindle forces produced by a tire
impacting large obstacles, Comput. Meth. Appl. Mech. Eng. (135) (1996) 15—34.
[10] I.M. Darnell, G.M. Hulbert, C.W. Mousseau, An efficient three-dimensional tire model for vehicle dynamics
simulations, Mech. Struct. Mach. 25 (1) (1997) 1—19.
[11] T.A. Laursen, J.C. Simo, A continuum-based finite element formulation for the implicit solution of multibody, large
deformation frictional contact problems, Int. J. Numer. Meth. Eng. 36 3451—3485.
[12] C.W. Gear, Numerical Initial Value Problems in Ordinary Differential Equations, Prentice-Hall, Englewood Cliffs,
NJ, 1971.
[13] M.A. Chace, Methods and experience in computer aided design of Large-displacement mechanical systems, in: E.J.
Haug (Ed.), Computer Aided Analysis and Optimization of Mechanical System Dynamics, NATO ASI Series, vol.
F9, Springer, Berlin, 1984.
[14] M. Padovano, Networking Applications on UNIX System V Release 4, Prentice-Hall, Englewood Cliffs, NJ, 1993.
[15] C.W., Mousseau, A numerical and experimental investigation into the force response of a tire impacting a large
obstacle, Ph.D. Dissertation, University of Michigan, 1994.
[16] J.P. Den Hartog, Mechanical Vibrations, McGraw-Hill, New York, 1934.

Vous aimerez peut-être aussi