Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Multibody System Dynamics

https://doi.org/10.1007/s11044-024-09984-2
RESEARCH

Real-time multibody simulation of vehicle wheel suspensions


of different topologies with elastokinematic properties

Jan-Lukas Archut1 · Burkhard Corves1

Received: 30 January 2024 / Accepted: 8 April 2024


© The Author(s) 2024

Abstract
Driving simulators are becoming increasingly important in vehicle development, as they
allow test drivers to evaluate the driving characteristics of a vehicle at an early stage of
the development process. Vehicle dynamics models used for such real-time applications are
usually simplified to achieve sufficiently low computation times, especially with respect to
the elastokinematic properties of the wheel suspension. This contribution introduces a real-
time capable multibody simulation model for wheel suspensions taking into account their
elastokinematics. The underlying data model allows the parametrization of different suspen-
sion topologies. Linear-implicit integration methods enable real-time capable integration of
the underlying numerically stiff equation of motion without major model simplifications.
It is shown how the computational effort of linear-implicit integration can be reduced by
explicitly taking the suspension topology into account. A versatile process for the analyti-
cal online-linearization of the equation of motion of arbitrary suspension topologies is pre-
sented. Both a double wishbone and a multilink suspension are simulated with very high
accuracy in real-time on a standard PC.

Keywords Vehicle wheel suspension · Elastokinematics · Real-time simulation · Numerical


stiffness · Linear-implicit integration

1 Introduction

The driving behavior of passenger vehicles is significantly influenced by the wheel sus-
pension. The wheel suspension guides the wheels relative to the vehicle body and defines
their position with respect to the road surface. This in turn affects the occurring tire forces
and thereby the vehicle’s motion. In addition to the purely kinematic guidance of the wheel,
wheel position changes occur under external loads due to the elastic properties of the suspen-
sion. This elasticity is mainly attributed to elastic bushings connecting the different compo-
nents of a wheel suspension with each other [1]. The design of these elastokinematic prop-
erties aims either to compensate the resulting wheel position changes or to convert them

 J.-L. Archut
archut@igmr.rwth-aachen.de

1 Institute of Mechanism Theory, Machine Dynamics and Robotics, RWTH Aachen University,
Eilfschornsteinstrasse 18, 52062 Aachen, Germany
J.-L. Archut, B. Corves

into desired wheel position changes to improve the vehicle’s driving behavior [1]. In auto-
motive engineering, a variety of wheel suspension concepts with different topologies has
emerged, such as McPherson, double wishbone or multilink suspension [2]. They exhibit
different freedom in the design of their kinematics and elastokinematics, but also different
complexity, costs, and packaging space requirements. This makes the development process
very challenging, as automotive engineers have to deal with different suspension topologies,
many design parameters, and also design objectives. Consequently, virtual vehicle dynamics
development is becoming increasingly important because it offers the potential for a more
targeted, cost-effective, and safer development process. Virtual development processes re-
quire multibody simulation models that allow the simulation of different suspension topolo-
gies with accurate and meaningful results within short computation times, e.g., to analyze
different design variants of a suspension. Moreover, not only objective design targets but
also the subjective perception of drivers play a crucial role in the vehicle dynamics design.
Hence, conducting test drives on driving simulators is crucial even in the initial phases of
development. However, vehicle dynamics models for such real-time applications are usually
simplified to meet the required computation times, including the elastokinematic properties
of the wheel suspension. In order to increase the validity of driving simulators and their ap-
plicability in the development process, real-time capable vehicle dynamics models without
such simplifications are desirable.
Real-time multibody simulation of vehicle wheel suspensions with elastic bushings is
challenging because of the high stiffness and damping properties of the bushings. Together
with the low inertia of the suspension bodies, high-frequency and/or strongly damped mo-
tions occur in the suspension system. In contrast, the dynamics of, e.g., the wheel travel
motion is comparatively slow. This means that the system features dynamics on different
time scales. From a numerical point of view, the system’s equation of motion has a high
numerical stiffness, making stable and real-time capable numerical integration challenging.
In literature, one can find many approaches for real-time multibody simulation of wheel
suspensions with special focus on their elastokinematics. This includes, for example, mod-
eling the wheel suspension as a macro joint [3], quasistatic approximation of the fast mo-
tions of the suspension bodies [4], or reducing the numerical stiffness through additional
stiffnesses and dampings in series to the bushings [5]. These approaches have in common
that the dynamics and/or the elastokinematics of the suspension are modified or simplified.
Alternatively, linear-implicit integration methods promise real-time numerical integration
of numerically stiff equations of motion without the need for major model simplifications.
Typical examples are the linear-implicit Euler method, Rosenbrock–Wanner or W-Methods
and the noniterative HHT-α method [6–8].
In this contribution, the authors present a multibody model for vehicle wheel suspen-
sions with different topologies. Based on a linear-implicit integration scheme, it allows for
real-time simulation without simplifications regarding their elastokinematic properties. With
the underlying data model, different suspension topologies can be set up and parametrized.
Elastic bushings are modeled as stiff, nonlinear force elements. The underlying numerically
stiff equation of motion is integrated with a linear-implicit L-stable real-time integrator with
two stages (LSRT2). Through exploiting the suspension topology, the computational effort
of the integration is significantly reduced. A calculation procedure is presented for analytical
online-linearization of the equation of motion, applicable to arbitrary suspension topologies.
The model is implemented in Matlab/Simulink. Using a double wishbone and a multilink
suspension as examples, computation times and accuracy of the model are evaluated for
dynamic load cases. The paper is structured as follows. In Sect. 2, the multibody systems
of the wheel suspensions are introduced, together with the underlying data model and their
RT-MBS of wheel suspensions with elastokinematic properties

numerically stiff equation of motion. In Sect. 3, challenges in real-time capable integration


are addressed and the linear-implicit integration method LSRT2 is described. Also, modifi-
cations for reducing the computational effort of the method are introduced. The calculation
procedure for the online-linearization of the equation of motion is explained in Sect. 4. Fur-
thermore, Sect. 5 presents the results in terms of accuracy and computation times. The paper
ends with the summary in Sect. 6.

2 Multibody systems of suspension systems

In the following, we consider two different suspension systems. This includes a double wish-
bone suspension as illustrated in Fig. 1 and a multilink suspension, typically employed at
the front and rear of mid-range passenger vehicles. Both suspensions feature extensive elas-
tokinematic properties through the use of elastic bushings.

2.1 Suspension topology

The double wishbone suspension features the topology shown in the topology diagram
Fig. 2 (left). The wheel is mounted on the wheel mount, which in turn is connected to the
vehicle body via the upper wishbone, the track rod, the trailing arm, and the lower wishbone.
The spring damper unit consists of the damper tube, the piston rod, and the spring which
is divided into an upper and a lower part. It is connected to the vehicle body and the lower
wishbone at both respective ends. An anti-roll bar connects the left- and right-hand side of
the suspension. It is divided into two rigid bodies with a rotational spring in between. On
its end, the anti-roll bar is attached to the damper tube via the anti-roll bar link. The bodies
are connected through elastic bushings to design the elastokinematic behavior of the sus-
pension. Furthermore, the force characteristics of the spring damper unit is modeled with
point-to-point (P2P) force elements. In addition, bump and rebound stops with highly pro-
gressive stiffness limit the wheel travel. All in all, the model consists of 24 bodies and 48
force elements, featuring a total of 144 degrees of freedom. The topology of the multilink
suspension, shown in Fig. 2 (right), differs from the double wishbone suspension. Here, the
wheel mount is guided through a total of five control arms. These are not directly connected

Fig. 1 Components of the double wishbone suspension (Color figure online)


J.-L. Archut, B. Corves

Fig. 2 Simplified topology diagrams of the double wishbone (left) and the multilink suspension (right); only
the left-hand side of the suspension systems is shown (Color figure online)

to the vehicle body, but to a subframe mounted on the vehicle body. The spring damper unit
with similar structure as in the double wishbone suspension is installed between the wheel
mount and the vehicle body. The anti-roll bar is attached to one of the control arms via
the anti-roll bar link. In total, the multilink suspension features a slightly higher number of
bodies (27) and force elements (63), and has 162 degrees of freedom.

2.2 Model elements and data model

The multibody systems of the two suspensions are composed of various types of model el-
ements, defined in the underlying data model. According to Fig. 3, the data model contains
all numerical parameters of the model elements. In addition, each model element is assigned
a unique index for identification. These indices are used to represent the specific suspension
topology in the data model. The first type of model elements are rigid bodies that represent
the different components of the suspension. They are parametrized by their mass and inertia
tensor and the position of the center of mass in the body coordinate system. Furthermore, the
initial pose of the body coordinate system in the global coordinate system is included in the
data model. A unique body index is assigned to each body in the suspension for identifica-
tion. Markers are local coordinate systems on the bodies as shown in Fig. 3. They are used to
track the motion of characteristic points and to apply forces and torques onto the bodies, e.g.,
of attached force elements. Each marker is defined by a unique marker index and uniquely
assigned to its corresponding body, whose body index is included in the marker properties.
To locate the marker on the body, its pose in the body coordinate system is included in the
data model. Bushings are force elements with generally nonlinear stiffness and damping
properties in all translational and rotational directions. They are used to model the elas-
tokinematic properties of the suspension, as they allow a certain flexibility in the connection
of the suspension components. A bushing is attached to two different markers (cf. Fig. 3)
and parametrized through characteristic curves that map the relative motion of the markers
to the force and torque output of the bushing. The mathematical formulation of the bushings
is based on the equations given in [9], which are extended to include relative motions and
torques in rotational direction and nonlinear characteristic curves. In accordance to typical
values given in [10], the stiffness values of the bushings are in the order of 105 ..107 N/m
RT-MBS of wheel suspensions with elastokinematic properties

Fig. 3 Data model for parametrizing a suspension system

in translational and 0..105 Nm/rad in rotational direction, depending on their specific loca-
tion in the suspension. Bushing force elements are also used to replace kinematic joints by
specifying high stiffness values in the directions to be constrained. This is the case, for ex-
ample, with the translational joints in the spring damper units. Each bushing force element is
uniquely identified through its bushing index. Additionally, the indices of its corresponding
markers are included in the bushing properties, as shown in Fig. 3. P2P elements are one-
dimensional force elements acting in the connection line between the two attached markers.
They are used to model the main spring and damper in the spring damper unit as well as
bump and rebound stops, parametrized through nonlinear characteristic curves. The use of
indices is the same as for bushing force elements. With this data model, arbitrary suspension
topologies can be modeled as illustrated in Fig. 3 (right). They are set up by adding further
model elements and connecting them to each other using the indices.

2.3 Equation of motion

The position and velocity state of a body l in the suspension system is described through the
body’s generalized coordinates rl and velocities vl defined as
0 T
rl = xl 0
yl 0
zl αz,l αy  ,l αx  ,l , (1a)
0 T
vl = vx,l 0
vy,l 0
vz,l B
ωx,l B
ωy,l B
ωz,l . (1b)

Here, 0 xl , 0 yl , 0 zl and 0 vx,l , 0 vy,l , 0 vz,l are translational displacements and velocities of
the body coordinate system in the global coordinate system. The rotation of the body is
described through Cardan angles αz,l , αy  ,l and αx  ,l with a yaw–pitch–roll current frame
rotation sequence and the angular velocities B ωx,l , B ωy,l , B ωz,l in the body coordinate sys-
tem. Singularities typical for Cardan angles are avoided by defining the body coordinate
system in such a way that there are no large rotations around the critical axis. Accordingly,
the equation of motion of a single body is formulated as

ṙl = Kl (rl )vl , (2a)


Ml v̇l = hl (r, v, u) = hs,l (r, v) − hω,l (rl , vl ) + hg,l (rl ) + hu,l (rl , vl , u), (2b)
J.-L. Archut, B. Corves

with the initial conditions rl,0 and vl,0 at time t = t0 . The kinematics matrix Kl (rl ) describes
the linear transformation of the generalized velocities to the time derivatives of the gener-
alized coordinates; Ml is the constant mass matrix of the body. The vector of generalized
forces hl consists of the body surface forces hs,l , gyroscopic forces hω,l , gravitational forces
hg,l , and input forces hu,l depending on external inputs to the suspension such as wheel
forces, denoted by the input vector u. The body surface forces arise from the bushing and
P2P force elements. As these elements always connect the body to other bodies in the sus-
pension, their force do not only depend on the generalized coordinates rl and velocities vl of
body l, but also on the values of the other bodies. Hence, hs,l is a function of the generalized
coordinates r and velocities v of the overall system. They are composed of the generalized
vectors of the individual bodies as
 T  T
r = rT1 rT2 . . . rTm , v = vT1 vT2 . . . vTm , (3)

where m represents the total number of bodies. Thus, the equation of motion of the overall
system is

ṙ = K(r)v, (4a)
Mv̇ = h(r, v, u), (4b)

with the mass matrix

M = diag(M1 , M2 , . . . , Mm ) , (5)

the kinematics matrix

K = diag(K1 , K2 , . . . , Km ) , (6)

and the composed vector of generalized forces


 T
h = hT1 hT2 . . . hTm . (7)

The equation of motion is a nonlinear ordinary differential equation featuring a high numeri-
cal stiffness because of the high stiffness and damping properties of the bushings. Numerical
stiffness can be quantified through the ratio maxi,j (|λi |/|λj |) of the eigenvalues λ of the lin-
earized equation of motion. A system is considered to be numerically stiff if the real parts
of all eigenvalues are negative and the ratio maxi,j (|λi |/|λj |) is much higher than one [11].
For the considered suspension systems, the ratio is higher than 105 , indicating a very high
numerical stiffness.

3 Linear-implicit integration of the equation of motion

Numerically stiff multibody systems are challenging in terms of real-time numerical in-
tegration. In real-time applications like driving simulators, where model outputs must be
provided at a fixed, predefined output rate, calculations must be executed in each time step
within the specified step size while ensuring numerical stability during the whole simulation.
On the one hand, implicit integration methods enable stable integration of numerically stiff
differential equations. However, they may require the iterative solution of nonlinear equation
systems, resulting in high computational effort that can even vary depending on the model’s
RT-MBS of wheel suspensions with elastokinematic properties

state and excitation. Consequently, numerical integration cannot be guaranteed to be com-


pleted in all integration steps within the step size specified by the real-time application. On
the other hand, explicit integration methods maintain a constant, a priori known computa-
tional effort per integration step, making them a popular choice for real-time applications.
However, they usually have poor stability properties so that integration of numerically stiff
differential equations may be unstable. In contrast, linear-implicit integration methods are a
suitable approach for stable and real-time capable integration of numerically stiff multibody
systems [6]. They are based on implicit integration methods but the equation of motion is
integrated in a linearized form. Thus, the nonlinear equation systems to be solved in im-
plicit integration methods are transformed to linear equation systems. Using direct solution
methods, these equation systems are solved with less computational effort which is even the
same in each integration step. Linear-implicit integration methods are therefore well suited
for real-time applications.

3.1 Integration method LSRT2

Several linear-implicit integration methods can be found in literature, some of which have
been compared by the authors in previous studies for the simulation of the double wishbone
suspension [12]. The L-stable real-time integrator with two stages LSRT2 was found to be
a good compromise between real-time capability and accuracy. It was originally proposed
for real-time dynamic testing of substructures subjected to dynamic loadings [13]. Based
on Rosenbrock schemes, which are a linear-implicit version of Runge–Kutta methods [8],
LSRT2 is a single-step integration method with s = 2 stages and high numerical damping
[13]. In this section, the method will be adapted for real-time capable time integration of the
suspension systems considered here. For this, the equation of motion (4a)–(4b) is linearized
according to

ṙ ≈ Klin v, (8a)
Mv̇ ≈ hlin + Jr,lin (r − rlin ) + Jv,lin (v − vlin ) + Ju,lin (u − ulin ), (8b)

with Klin = K(rlin ), hlin = h(rlin , vlin , ulin ) and the Jacobians

∂h ∂h ∂h
Jr,lin = , Jv,lin = , Ju,lin = (9)
∂r ∂v ∂u

in the linearization state rlin , vlin , and ulin . Based on the known generalized coordinates rk
and velocities vk in time step k, the values in the next time step k +1 are calculated according
to


s
rk+1 = rk + t (bi rk,i ), (10a)
i=1


s
vk+1 = vk + t (bi vk,i ), (10b)
i=1
J.-L. Archut, B. Corves

with the constant integration stepsize t = tk+1 − tk . For calculating vk,i in the stage i, the
following linear equation system has to be solved:
 
M − tγii Jv,lin − (tγii )2 Jr,lin Klin vk,i
= h(r̃k,i , ṽk,i , ũk,i ) + tγii Ju,lin uk,i
i−1 
   T 
+ t Jr,lin Jv,lin Ju,lin γij (rk,j )T (vk,j )T (uk,j )T (11)
j =1
⎡ ⎤

i−1
 
+ tγii Jr,lin Klin ⎣ṽk,i + t γij v k,j ⎦.
j =1

Here, uk,j is defined as


1 k+1,j
uk,j = (u − uk ) . (12)
t

Afterwards, rk,i is calculated according to


⎡ ⎤

i
r k,i
= Klin ⎣ṽ k,i
+ t k,j ⎦
(γij v ) . (13)
j =1

The stage vectors r̃k,i , ṽk,i , and ũk,i are defined as


i−1
r̃k,i = rk + t (aij rk,j ), (14a)
j =1


i−1
ṽk,i = vk + t (aij vk,j ), (14b)
j =1

ũk,i = u(t k + ci t) . (14c)

The parameters of the method s, bi , ci , aij , and γij are given in Table 1. To reduce the
computational effort of the method, the above equations are composed and simplified. Here,
it is exploited that the stage vectors r̃k,i , ṽk,i , and ũk,i do not need to be explicitly calculated,
and the vector h(r̃k,i , ṽk,i , ũk,i ) is approximated linearly similar to the right-hand side of
equation (8b). With that, the linear equation system (11) simplifies to
 
M − tγii Jv,lin − (tγii )2 Jr,lin Klin vk,i
= hlin + Jr,lin (r̄k,i − rlin + tγii Klin v̄k,i ) (15)
+ Jv,lin (v̄ k,i
− vlin ) + Ju,lin (ū k,i
− ulin ),

with the vectors


i−1
r̄k,i = rk + t ((aij + γij )rk,j ), (16a)
j =1
RT-MBS of wheel suspensions with elastokinematic properties

Table 1 Parameters of the


method LSRT2 [13] s b1 b2 c1 c2 a21 γ11 = γ22 γ21
√ √
2 0 1 0 1
2
1
2 1 − 22 −(1 − 22 )


i−1
v̄k,i = vk + t ((aij + γij )vk,j ), (16b)
j =1


i
ūk,i = u(t k + ci t) + t (γij uk,j ) . (16c)
j =1

Similarly, (13) simplifies to


 
rk,i = Klin v̄k,i + tγii vk,i . (17)

These simplifications reduce the number of required matrix vector multiplications with the
Jacobians, thus decreasing the computational effort.

3.2 Exploiting the suspension topology in linear-implicit integration

In each stage i = 1, . . . , s of an integration step, a linear equation system

Ax = b (18)

with the coefficient matrix


 
A = M − tγii Jv,lin − (tγii )2 Jr,lin Klin (19)

must be solved, cf. (15). The solution is done by means of LU decomposition and for-
ward/backward substitution. With γ11 = γ22 , the coefficient matrix remains constant for all
stages of an integration step. In addition to that, the coefficient matrix only changes when
the Jacobians Jr,lin and Jv,lin , as well as the kinematics matrix Klin , are updated. Hence, LU
decomposition must be performed only after an update of the linearization.
By exploiting the topology of a specific suspension, the computational effort for solving
the linear equation system can be reduced without sacrificing the accuracy of the solution.
As the coefficient matrix is a function of the mass matrix, the kinematics matrix and the
Jacobians of the suspension system, its structure is strongly connected to the suspension
topology. The coefficient matrices for the double wishbone and the multilink suspension are
block matrices with the colored blocks as shown in Fig. 4. The colors of the main-diagonal
blocks refer to the colors in the topology diagrams in Fig. 2. The bodies that connect the left
and right sides of the suspensions are colored blue. These are the anti-roll bar and, in the
case of the multilink suspension, also the subframe. The subsystem wheel, wheel mount and
control arms is colored in red and the spring damper subsystem is colored in yellow. The
nonzero elements outside the matrix diagonal result from the connections between these
subsystems. In the double wishbone suspension, the anti-roll bar is connected to the spring
damper unit on both sides. The subsystem wheel, wheel mount and control arms is also
connected to the spring damper unit, but not to the anti-roll bar. Accordingly, there are only
nonzero blocks on the main and on the lower and upper diagonal of the block matrix. The
same applies for the multilink suspension.
J.-L. Archut, B. Corves

Fig. 4 Structure of the coefficient matrix (19) for the double wishbone (left) and multilink suspension (right)
(Color figure online)

This matrix structure can be utilized to solve the linear equation system (18) more effi-
ciently. To this end, the system is written with the coefficient matrix in block matrix form:
⎛ ⎞⎛ ⎞ ⎛ ⎞
A11 A12 0 0 0 x1 b1
⎜ A21 A22 A23 0 0 ⎟ ⎜ x2 ⎟ ⎜ b2 ⎟
⎜ ⎟⎜ ⎟ ⎜ ⎟
⎜ 0 0 ⎟ ⎜ ⎟ ⎜ ⎟
⎜ A32 A33 A34 ⎟ ⎜ x3 ⎟ = ⎜ b3 ⎟ . (20)
⎝ 0 0 A43 A44 A45 ⎠ ⎝ x4 ⎠ ⎝ b4 ⎠
0 0 0 A54 A55 x5 b5

This linear equation system is then divided into several smaller systems that are solved
subsequently. Firstly, the linear equation system

Ã33 x3 = b̃3 (21)

based on the coefficient matrices

Ã33 = A33 − A32 Ã−1 −1


22 A23 − A34 Ã44 A43 , (22a)

Ã22 = A22 − A21 A−1


11 A12 , (22b)

Ã44 = A44 − A45 A−1


55 A54 , (22c)

and the right-hand side vectors

b̃3 = b3 − A32 Ã−1 −1


22 b̃2 − A34 Ã44 b̃4 , (23a)

b̃2 = b2 − A21 A−1


11 b1 , (23b)

b̃4 = b4 − A45 A−1


55 b5 , (23c)

is solved. Afterwards, the equation systems

Ã22 x2 = b̃2 − A23 x3 , Ã44 x4 = b̃4 − A43 x3 , (24a)


A11 x1 = b1 − A12 x2 , A55 x5 = b5 − A54 x4 (24b)
RT-MBS of wheel suspensions with elastokinematic properties

are solved. Savings of computation time result from the fact that several small linear equa-
tion systems are solved instead of one large equation system, requiring less computational
effort for LU decomposition and forward/backward substitution. Furthermore, only the
nonzero elements in the matrices Aij (i = j ) are considered in the calculations of equations
(22a)–(22c)–(24a)–(24b). It should be noted that the partitioning of the coefficient matrix
according to Fig. 4 may not be the best in terms of computational efficiency. However, it is
closely related to the suspension topology and therefore straightforward to formulate.

4 Online-linearization of the equation of motion


For time integration with the LSRT2 integration method, the equation of motion has to be
linearized in a certain linearization state rlin , vlin , ulin to provide the generalized forces hlin ,
the Jacobians Jr,lin , Jv,lin , Ju,lin , and the kinematics matrix Klin . There are several possible
choices for the linearization state, which affect computational effort and accuracy of the
method. Linearization could be performed only once at the beginning of the simulation in
the initial configuration. However, as the suspension systems considered here are strongly
nonlinear, the linearization has to be updated online during the simulation. The update is
possible in each integration step or only after a few integration steps with a fixed lineariza-
tion step size tint that is an integer multiple of the integration step size t . Since both the
integration and linearization step size influence accuracy and computational effort, a suitable
choice must be made as shown in Sect. 5.
To meet the real-time requirements of an online-linearization, the equation of motion
is linearized analytically. Furthermore, a calculation procedure suitable for any arbitrary
suspension topology, as for example the double wishbone and the multilink suspension,
is desired. Therefore, the generalized forces hlin , the Jacobians Jr,lin , Jv,lin , Ju,lin , and the
kinematics matrix Klin are calculated following a procedure based on the data model (cf.
Sect. 2.2) as shown in Fig. 5. The procedure starts with the calculation of the body kine-
matics. The kinematics of each body l is calculated separately based on its generalized co-
ordinates rl and velocities vl , which are extracted from the generalized coordinates rlin and

Fig. 5 Procedure for analytical online-linearization of the equation of motion


J.-L. Archut, B. Corves

velocities vlin of the overall system. With that, all kinematic variables together with their
partial derivatives required for the subsequent calculation steps are calculated. Additionally,
the kinematics matrix Klin of the overall system is composed. In the next step, the marker
kinematics is calculated. To calculate the kinematics of a specific marker, it is necessary to
know the kinematic variables of the body associated with it. These variables are accessed
through the body index that is stored in the marker properties. Again, not only the kinematic
variables of the markers, but also their partial derivatives are calculated, for each marker
separately. These values serve as input for the next calculation step, in which the force el-
ements (bushings, P2P) are calculated. The kinematic variables of the markers to which a
force element is attached are read based on their indices stored in the force element prop-
erties. Then, the relative motion of the markers is calculated and the forces and torques of
the force element are determined based on its characteristic curves. Again, the required par-
tial derivatives are calculated. This procedure is repeated for all force elements. In the next
step, the generalized surface forces hs,l are computed for each individual body sequentially,
considering all force elements that are attached to the body. These force elements, as well as
kinematic variables of the markers and the body, are accessed through the indices stored in
the data model. Also, the derivatives of the body surface forces are computed by combining
the previously calculated partial derivatives using the chain rule. Furthermore, the vector of
generalized forces hlin and the Jacobians Jr,lin and Jv,lin are initialized and filled with the
corresponding values of the body surface forces. In the next calculation steps, the body gy-
roscopic forces hω,l and gravitational forces hg,l are calculated based on the corresponding
body kinematics. Their quantities and derivatives are added to the generalized forces and
Jacobians of the overall system. The last step of the linearization procedure is the calcula-
tion of the body input forces hu,l based on the inputs ulin . They are added to the generalized
forces of the overall system. Moreover, the Jacobian Ju,lin is set up. This procedure based on
the data model allows for linearizing arbitrary suspension systems as their topology is auto-
matically considered through the indices stored in the data model. Additionally, analytical
calculation enables updating the linearization online during real-time simulation.

5 Results

In this section, the results in terms of accuracy for dynamic load cases and the required
computation times are presented.

5.1 Simulation of dynamic load cases

To demonstrate the accuracy of the model and the integration method, dynamic load cases
were performed with both suspension systems. The first load case (LC1) is a force step input
of the longitudinal wheel force according to

0 N, t < 5.0 s,
Fx (t) = (25)
−2500 N, t ≥ 5.0 s,

where the positive direction points in driving direction. The second load case (LC2) is a
force frequency sweep of the longitudinal wheel force according to

−2500 N, t < 5.0 s,
Fx (t) = (26)
−2500 N + 500 N · sin( 2π
s2
(t − 5.0 s)2 ), 5.0 ≤ t ≤ 20.0 s,
RT-MBS of wheel suspensions with elastokinematic properties

with a maximum frequency of 30 Hz. The upward vertical wheel force is Fz = 5000 N in
both load cases. All wheel forces are applied on both wheels at the respective wheel center.
Integration was performed with the LSRT2 method with different integration step sizes t
and linearization step sizes tlin to evaluate their influence on the simulation results. The
results were compared to reference results. To this end, the double wishbone suspension
system was also implemented in the commercial multibody software Simpack and reference
simulations were performed with the Simpack solver SODASRT 2, which is an implicit
multistep solver suitable for stiff differential equations [14]. For the multilink suspension,
reference results were obtained in Matlab/Simulink. The Matlab/Simulink solver ode23tb
was used as reference for LC1. ode23tb is an implicit solver for stiff differential equations
based on a Runge–Kutta formula with a trapezoidal rule step as first stage and a backward
differentiation as second stage [15]. For LC2, the LSRT2 method was used with a small
step size of 10−3 ms. The reference simulations in Simpack, as well as in Matlab/Simulink,
required computation times far from real-time, but serve as a reference for evaluating the
LSRT2 method due to their accuracy. For all simulations, the normalized root mean square

Fig. 6 Longitudinal wheel displacement and bushing force of the double wishbone and multilink suspen-
sion obtained with LSRT2 with different integration step sizes t (update of the linearization before each
integration step, t = tlin ) for LC1 (Color figure online)
J.-L. Archut, B. Corves

error to the reference results was calculated according to


  
i=1 (yi − yi,ref )
n 2
1
ey = , (27)
ȳref n

where yi and yi,ref are the value of the signal to be evaluated and the reference signal at
sample i, respectively, and ȳref is the mean value of the reference signal; n refers to the total
number of samples, with all signals resampled to a sampling time of 1.0 ms.
Figure 6 shows the time plots of the longitudinal wheel displacement and the force of a
bushing for LC1. After the wheel force step, weakly damped oscillations of the longitudinal
wheel displacement and the bushing force occur. According to the elastokinematic prop-
erties of the suspensions, the wheel displacement tends towards a static equilibrium with
a displacement of a few millimeters. Comparing the simulations with different integration
step sizes, stable simulation is possible with step size t = 5.0 ms, but with high devia-
tions from the reference results, which are clearly visible in the time plots. In contrast, the
deviations obtained with smaller step sizes are almost within the line width. The results for
LC2 are shown in Fig. 7. The plots show the envelopes of the oscillations that result from
the force frequency sweep excitation. Similarly to LC1, high deviations occur for high step
sizes. The amplitude gain at approximately t = 15 s cannot be reproduced accurately with
step sizes t = 5.0 ms and t = 2.0 ms. For smaller step sizes, the deviations are again
within the line width. A more detailed view on the accuracy for both load cases is given

Fig. 7 Envelope of longitudinal wheel displacement and bushing force of the double wishbone and multilink
suspension obtained with LSRT2 with different integration step sizes t (update of the linearization before
each integration step, t = tlin ) for LC2 (Color figure online)
RT-MBS of wheel suspensions with elastokinematic properties

Fig. 8 Error values of the wheel displacement ex and the bushing force ef for LC1 (Color figure online)

Fig. 9 Error values of the wheel displacement ex and the bushing force ef for LC2 (Color figure online)

in Figs. 8 and 9. Through reducing the integration step size and updating the linearization
before each integration step (t = tlin ), the deviation from the reference results is reduced
with a convergence to a remaining deviation. Results with an error in the range of 1% can al-
ready be obtained with a step size of t = 1.0 ms. In contrast, reducing the integration step
size while keeping the linearization step size constant does not significantly reduce the error.
This indicates the high nonlinearity of the considered suspension systems, which requires
frequent updating of the linearization during simulation.

5.2 Computation times

To evaluate the required computation times, the Matlab/Simulink model was compiled with
the Simulink Coder and C code was generated [16]. The Intel oneAPI Library containing
J.-L. Archut, B. Corves

Table 2 Obtained real-time


factors with integration step size Double Multilink
t = 1.0 ms (with and without wishbone suspension
exploiting the suspension suspension
topology according to Sect. 3.2)
Without exploiting the suspension topology 0.552 0.761
With exploiting the suspension topology 0.494 0.676

implementations of Basic Linear Algebra Subprograms (BLAS) was incorporated in the


compilation process [17]. Thus, Simulink Coder generates BLAS calls to accelerate linear
algebra operations. The generated code was executed on a standard desktop computer with
an Intel Core i7-3770 CPU at 3.4 GHz. The obtained real-time factors, defined as the ratio
of the required computation time to the specified duration of a simulation, are shown in
Table 2 for integration step size t = 1.0 ms (linearization before each integration step, i.e.,
t = tlin ). It can be seen that all real-time factors are less than one, which means that
all simulations are real-time capable. The real-time factors of the multilink suspension are
higher than those of the double wishbone suspension because of its higher number of model
elements and degrees of freedom. By exploiting the suspension topology, computation times
can be reduced by up to almost 12%. This results in real-time factors of 0.494 and 0.676 for
the double wishbone and the multilink suspension, respectively, which are well below one.
Considering the error of about 1% obtained with the integration step size t = 1.0 ms, it can
be concluded that simulation results with very good accuracy can be provided in real-time
for the considered suspension systems.

6 Conclusion
This contribution demonstrates the real-time capable simulation of numerically stiff multi-
body systems, focusing on vehicle wheel suspensions. Here, high numerical stiffness results
from elastic bushings connecting the suspension bodies. Real-time capable simulation with-
out model simplifications is achieved using the linear-implicit integration method LSRT2,
transforming the nonlinear equation system to be solved for implicit integration into a lin-
ear equation system. It is shown how the suspension topology can be exploited to reduce the
computational effort for solving the linear equation system. With the underlying data model,
wheel suspensions with arbitrary topologies can be parameterized. Furthermore, an analyt-
ical online-linearization procedure allows to linearize the equation of motion of arbitrary
suspension systems during real-time simulation. This requires that the derivatives of the dif-
ferent model elements in the suspensions can be derived analytically. The applicability of
the model is demonstrated using two suspensions systems with different topologies, a double
wishbone and a multilink suspension. With real-time factors of 0.494 and 0.676, both sus-
pension systems can be simulated in real-time with very high accuracy (error of only about
1%) for the considered dynamic load cases. For application in human/hardware-in-the-loop
applications such as driving simulators, the suspension systems have to be integrated into
a full vehicle model. Assuming that a vehicle consists of different subsystems that can be
simulated in parallel, real-time capability is also expected for such a full vehicle model.

Author contributions J.-L. A. and B.C. made the concept of the work. J.-L. A. developed and implemented
the simulation model and analyzed the results. J.-L. A. prepared the original draft of the manuscript. B.C.
reviewed and edited the manuscript. All authors approved the version of the manuscript to be published.

Funding Open Access funding enabled and organized by Projekt DEAL.


RT-MBS of wheel suspensions with elastokinematic properties

Data availability No datasets were generated or analysed during the current study.

Declarations

Competing interests The authors declare no competing interests.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which
permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence,
and indicate if changes were made. The images or other third party material in this article are included in the
article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is
not included in the article’s Creative Commons licence and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

References
1. Matschinsky, W.: Road Vehicle Suspensions. Professional Engineering Publ., London (1997). ISBN 978-
1860582028
2. Trzesniowski, M.: Suspension System. Springer, Heidelberg (2023). ISBN 978-3-658-39846-0
3. Eichberger, A., Rulka, W.: Process save reduction by macro joint approach: the key to real time
and efficient vehicle simulation. Veh. Syst. Dyn. 41(5), 401–413 (2004). https://doi.org/10.1080/
00423110412331300354
4. Burgermeister, B., Arnold, M., Eichberger, A.: Smooth velocity approximation for constrained systems
in real-time simulation. Multibody Syst. Dyn. 26(1), 1–14 (2011). https://doi.org/10.1007/s11044-011-
9243-1
5. Kracht, F.E.: Modellbildung und Simulation der Dynamik und Elastokinematik von Radaufhängungen
für Echtzeitanwendungen. PhD Thesis, DuEPublico: Duisburg-Essen Publications online, University of
Duisburg-Essen, Germany (2020). https://doi.org/10.17185/duepublico/71445
6. Arnold, M., Burgermeister, B., Eichberger, A.: Linearly implicit time integration methods in real-time
applications: DAEs and stiff ODEs. Multibody Syst. Dyn. 17, 99–117 (2007). https://doi.org/10.1007/
s11044-007-9036-8
7. Kim, M., Song, H., Kim, S.S.: A non-iterative implicit integration method using a HHT-α integrator for
real-time analysis of multibody systems. J. Mech. Sci. Technol. 33(3), 1087–1096 (2019). https://doi.
org/10.1007/s12206-019-0208-2
8. Rang, J.: Improved traditional Rosenbrock–Wanner methods for stiff ODEs and DAEs. J. Comput. Appl.
Math. 286, 128–144 (2015). https://doi.org/10.1016/j.cam.2015.03.010
9. Rill, G., Schaeffer, T., Borchsenius, F.: Grundlagen und computergerechte Methodik der Mehrkörper-
simulation. Springer, Wiesbaden (2020). ISBN 978-3-658-28912-6
10. Heißing, B., Ersoy, M.: Chassis Handbook: Fundamentals, Driving Dynamics, Components, Mechatron-
ics, Perspectives. Vieweg+Teubner, Wiesbaden (2010). ISBN 978-3-8348-9789-3
11. Dahmen, W., Reusken, A.: Numerik für Ingenieure und Naturwissenschaftler, 2nd edn. Springer, Berlin
(2008). ISBN 978-3-540-76493-9
12. Archut, J.L., Corves, B.: Comparison of integration methods for real-time capable multibody simulation
of an elastokinematic wheel suspension system. In: Joint International Conference on Multibody System
Dynamics, New Delhi, India (2022)
13. Bursi, O.S., Gonzalez-Buelga, A., Vulcan, L., Neild, S.A., Wagg, D.J.: Novel coupling Rosenbrock-
based algorithms for real-time dynamic substructure testing. Earthq. Eng. Struct. Dyn. 37(3), 339–360
(2008). https://doi.org/10.1002/eqe.757
14. Dassault Systemes Simulia Corp.: Simpack Assistant 2023: SODASRT 2 (2023)
15. MathWorks: Solve stiff differential equations – trapezoidal rule + backward differentiation formula –
MATLAB ode23tb (19.08.2022). https://de.mathworks.com/help/matlab/ref/ode23tb.html
16. MathWorks: Simulink Coder Documentation (19.08.2022). https://de.mathworks.com/help/rtw/
17. Intel: Accelerate Fast Math with Intel oneAPI Math Kernel Library (19.08.2022). https://software.intel.
com/oneapi/onemkl

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

You might also like