optimal motorcycle braking

Embed Size (px)

Citation preview

  • 8/18/2019 optimal motorcycle braking

    1/14

    Control Engineering Practice 16 (2008) 644–657

    On optimal motorcycle braking

    Matteo Cornoa, Sergio Matteo Savaresia,, Mara Tanellia, Luca Fabbrib

    aDipartimento di Elettronica e Informazione, Politecnico di Milano, Piazza L. da Vinci, 32, 20133 Milano, ItalybPiaggio Group—Aprilia Brand, Via Galileo Galilei, 1, 30033 Noale, Venice, Italy

    Received 21 March 2007; accepted 1 August 2007

    Available online 21 September 2007

    Abstract

    The optimal braking strategy in a high-performance motorbike is discussed. First, the control strategy using the front brake only isanalyzed, highlighting the role of aerodynamics and studying how to select and modify the control objective during braking. The

    importance of the brake modulation in the very first part of the braking maneuver is also discussed. The role played by the rear brake is

    then analyzed. Finally, the attention is shifted to the damping ratio of the front suspension, which is probably the single motorbike

    parameter with the largest impact on vehicle dynamics during braking. A possible policy for semi-active suspension control during

    braking is sketched.

    r 2007 Elsevier Ltd. All rights reserved.

    Keywords:  Motorbike; Brake control; Semi-active suspension; Vehicle dynamics; Automotive systems

    1. Introduction and motivation

    In this work a simulation-based study of the brakingmaneuver in high-performance (racing) motorcycles is

    presented.

    The starting point of this work is the analysis of a typical

    hard-braking maneuver performed by a professional

    driver. An example is given in  Fig. 1, where the measured

    front-wheel speed, front-brake pressure and suspensions

    elongation are displayed (the measurements are made on

    the MY05 Aprilia RSV1000 Factory). During this braking

    maneuver, the driver makes no use of the rear brake.

    Notice that the brake pressure has a ramp (of about

    250 ms), followed by a small brake release (probably due to

    an over-slip phenomenon sensed by the driver); then thebrake pressure is kept almost constant until the end of 

    the braking maneuver. By inspecting the elongation of the

    front and rear suspensions, it is interesting to observe that

    the suspension hard-limit is reached in both cases; notice

    that this means that the rear-wheel is at the contact limit;

    apparently, this may induce the driver to avoid turning the

    vehicle before the braking maneuver is ended.

    Among pilots and race engineers it is common opinion

    that the braking phase is the most critical and sensitive

    maneuver. The ability of a driver to achieve an ‘‘optimalbraking’’ can make the difference on the lap-time. Even few

    milliseconds per braking hence can be crucial. In this work,

    optimality simply means minimum time to decelerate the

    motorbike from the initial speed to a target speed. A

    broader range of performance parameters will be discussed

    in Section 4, where the sensitivity to the front-suspension

    damping will be analyzed.

    The objective of this paper is to deeply analyze a single

    braking maneuver, trying to understand what the optimal

    maneuver should be, and the main parameters which can

    influence it.

    The analysis is developed in the following setting:

     A pure braking maneuver is assumed, on a straight line(no lateral forces are engaged).

      An ideal driver is assumed; the concept of the idealdriver can be recast into an automatic closed-loop

    control system, following a reference signal (e.g., a

    target slip or a target load). The design of the reference

    signal is clearly a key issue for the optimality (‘‘ideality’’)

    of the driver.

    ARTICLE IN PRESS

    www.elsevier.com/locate/conengprac

    0967-0661/$- see front matter r 2007 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.conengprac.2007.08.001

    Corresponding author. Tel.: +39 02 23993545; fax: +39 02 23993412.

    E-mail address:  [email protected] (S.M. Savaresi).

    http://www.elsevier.com/locate/conengprachttp://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.conengprac.2007.08.001mailto:[email protected]:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.conengprac.2007.08.001http://www.elsevier.com/locate/conengprac

  • 8/18/2019 optimal motorcycle braking

    2/14

    The main challenge in the development of the ‘‘optimal

    maneuver’’ is due to the fact that a complete analytical

    model of a motorcycle is very complex, and it can hardly beused for the direct closed-form calculation of the optimal

    solution. On the other hand, reduced order models can

    indeed offer analytic solutions, but they cannot be really

    useful to find the optimum for the real target vehicle. This

    classical dilemma is solved according to the objective and

    the perspective of the specific research work; in this paper,

    since the objective is to analyze the very details of a braking

    maneuver, the simulator-based approach has been the

    natural choice (Fig. 2).

    The analysis developed in this work is based on a full-

    fledged motorcycle simulator (the Mechanical Simulation

    Corp.   BikeSims simulation environment, based on the

    AutoSim symbolic multi-body software (Sharp, Evangelou,

    &  Limebeer, 2004, 2005), which takes into account all the

    motorcycle dynamics, and accurately models the road–tire

    interaction forces (Sharp et al., 2004). See also   Cossalter

    and Lot (2002),  Donida, Ferretti, Savaresi, Schiavo, and

    Tanelli (2006), Frezza and Beghi (2005), Lin, Chyuan-Yow,

    and Tsai-Wen (2006),   Sayers (1999), and   Sharp and

    Limebeer (2001)   for other state-of-the-art simulation

    environments suitable for this kind of analysis. The

    complete set of parameters used in the simulator is listed

    in Appendix A.

    Solving the problem of the ‘‘optimal’’ maneuver is very

    attractive, since it allows to decouple the intrinsic

    performance of the vehicle from the driver behavior. In

    2-wheel vehicles, these two aspects are so strictly inter-

    leaved that it is hard to predict the effect of a parameter

    change on the vehicle. This is a well-known problem, which

    sometimes has the picturesque effect of transforming the

    tuning of a racing motorbike from a rigorous model-based

    procedure into a sort of fine (or ‘‘magic’’) art.

    ARTICLE IN PRESS

    0 1 2

    0

    100

    200

       [   k  m   /   h   ]

    Front-wheel speed

    0 1 2

    0

    5

    10

    15

    20

       [   b  a  r   ]

    Front brake pressure

    0 120

    10

    20

    30

    40

       [  m  m   ]

    Rear suspension extension

    0 1 2

    0

    50

    100

    time [s]

       [  m  m   ]

    Front suspension extension

    START BRAKING

    250ms

    Temporary brake release Steady-state braking pressure

    Rebound hard-limit

    Compression hard-limit

    Fig. 1. Typical hard-braking maneuver (measured on the MY05 Aprilia RSV1000 Factory). The convention used for the suspension extension is that the

    maximum value is taken at full compression, the minimum at full extension.

    F  z f F  zr  F  x r F  x  f 

    wheelbase  p

    b

    h

    Fig. 2. Notation and GUI of the hypersport-class motorcycle used in the

    SW simulator Bikesim.

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657    645

  • 8/18/2019 optimal motorcycle braking

    3/14

    The optimal-maneuver problem is a very challenging

    task, which has been studied in the last decade by different

    research groups, from different perspectives (see e.g.,

    Cossalter, Doria,  & Lot, 1999;  Cossalter, Lot,  & Maggio,

    2002;  Frezza, Beghi,   &   Saccon, 2004;   Hauser   &   Saccon,

    2006; Hauser, Saccon, & Frezza, 2004; Limebeer, Sharp, &

    Evangelou, 2001; Sharp, 1971, 2001 for an overview of theliterature on this topic). The specific problem of optimal

    braking has been considered in Cossalter, Lot, and Maggio

    (2004), where the focus was on steady-state conditions

    during a bend. Another paper dealing with the braking

    maneuver is   Cossalter, Doria, and Lot (2000), where the

    design of the suspension for a scooter is discussed. The

    present work focuses on the dynamic aspects of a pure

    braking maneuver on a straight line.

    In this paper the problem of the optimal braking

    maneuver is discussed following this path:

      Section 2. As first step, it is assumed that only thefront brake is used (this is quite common, even inracing). Under this assumption the optimal control

    target is discussed; first this analysis is made by

    considering quasi-stationary conditions; then the fast

    transient of the very first part of the braking maneuver is

    analyzed.

     Section 3. The role and the benefits of rear braking arediscussed.

      Section 4. Since the front-suspension damping is verycritical for the overall braking performance, the

    influence of this parameter on the braking maneuver is

    analyzed.

    The approach used in this paper differs from the classical

    automatic-control design path (system modeling, para-

    meter identification, control algorithms design, perfor-

    mance evaluation). A mix of analysis of vehicle dynamics

    and closed-loop control architectures is proposed; as a

    matter of fact the primary goal of this paper is not the

    complete design of an automatic braking control system,

    but the understanding of the main dynamic phenomena

    and reference signals to be considered in the design of an

    optimal braking control strategy.

    2. Optimal front-braking

    In a high-performance motorbike it is customary to use

    only the front brake during a hard-braking maneuver.

    Since no front–rear coupling phenomena must be taken

    into account, and almost all the vehicle load is transferred

    onto the front wheel, the braking dynamics can be well-

    approximated with a single-corner model (Cossalter, 2002;

    Savaresi, Tanelli,   &   Cantoni, 2007;  Solyom, Rantzer,   &

    Lu ¨ demann, 2004):

    J   _o ¼  rF x   T ;

    m_

    v ¼ F x;(   (1)

    where o  is the angular speed of the wheel (it is expressed in

    rad/s;  o40 is assumed),  v   is the longitudinal speed of the

    vehicle body,  T  the braking torque which plays the role of 

    control/input variable,   F x   the longitudinal road–tire con-

    tact force, J , m  and  r  are the rotational inertia of the wheel,

    the corner mass and the wheel radius, respectively.

    The dynamic behavior of the system is hidden in theexpression of   F x, which depends on the state variables   v

    and   o. The most general expression of   F x   is quite

    complicated, since   F x   depends on a large number of 

    features of the road, tire and suspension; however, it can be

    well-approximated as   F x  ¼ F zmðl; btÞ, where   F z   is thevertical force at the tire–road contact point and   l   is

    the longitudinal slip, given by   l ¼ ðor  vÞ=v ¼  ro=v  1(the sign convention of negative slip when braking is used);

    bt is the side-slip angle of the wheel. Since it is assumed that

    the braking maneuver is performed along a straight line

    (bt ¼  0), the dependence of  m( ) on  bt  will be omitted.

    The function m(l) is called friction coefficient. It can be

    modeled using many empirical or semi-empirical para-

    metric models, among which the most famous is probably

    the Pacejka ‘‘magic’’ formula (Pacejka, 2002;   Savaresi,

    Tanelli, Langthaler,   &   Del Re, 2006). The curve   m(l) is

    characterized by a peak, which is typically around

    lE0.15; this non-monotonic characteristic is responsible

    for the well-known fact that (for constant braking torque

    values) the equilibria associated to slip values beyond the

    peak of  m(l) are unstable. In Fig. 3, the road–wheel friction

    curve of the front tire used in the simulator is depicted, for

    different vertical contact forces  F z. Notice that, at different

    vertical loads, this curve slightly changes. In the rest of the

    paper it is assumed that the maximum friction forceis achieved at   ¯ l ¼ 0:12. This is rigorously true if F z ¼  2000N ; since the total mass of the vehicle and driver

    is about 270 kg—see Appendix A—this vertical load

    corresponds to a significant load transfer onto the front

    axle. Notice that for higher vertical loads the peak position

    slightly changes, but by keeping the target slip at  ¯ l ¼ 0:12the loss of performance is negligible (see details on Fig. 3).

    The shape of the friction curve provides a seemingly simple

    solution to the optimal braking control target: in order to

    maximize the longitudinal force, the longitudinal slip l should

    be simply regulated at the peak value of the friction curve,

    whatever is the vehicle speed and the vertical load.

    Fig. 4  shows the results of a braking maneuver, using

    only the front brake and a target-slip control policy. The

    results displayed in Fig. 4 are obtained using a well-tuned

    PID slip regulator, applied to the Bikesim simulator (see

    Buckholtz, 2002; Johansen, Petersen, Kalkkuhl, & Lu ¨ demann,

    2003 for details on the design of such a slip controller). The

    results displayed in   Fig. 4   clearly show that—unfortu-

    nately—in a motorcycle the target-slip policy cannot be

    implemented, since the maximum braking torque can be

    higher than the overturn torque. This is a very peculiar

    feature of motorbikes (this problem does not exist in racing

    cars); it is clearly due to the high ratio between the height

    of the center of mass and the wheelbase (see  Fig. 2).

    ARTICLE IN PRESS

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657 646

  • 8/18/2019 optimal motorcycle braking

    4/14

    By inspecting Fig. 4, it is worth pointing out that:

     the braking maneuver starts at a very high longitudinalspeed (300 km/h);

     the PID-based automatic-control system is able to trackvery accurately (with both dynamic and static precision)

    the reference slip target  ð¯ l ¼ 0:12Þ;  the vertical load at the rear wheel ðF zr Þ rapidly decreases,

    while the pitch angle (y) increases; after 2.2 s from the

    braking start, at the speed of about 200 km/h and at the

    pitch angle of about 101, the rear wheel starts losing

    contact with the road;

      the front-wheel target-slip is kept at   ¯ l ¼ 0:12 evenafter the loss of contact; this causes an overturn of the

    vehicle (also called a ‘‘stoppie’’ in the motorcycle

     jargon—see photo in Fig. 4).

    In order to better understand the reason of this

    phenomenon, a simplified model of the in-plane dynamics

    of the motorcycle can be used. Note that this model does

    ARTICLE IN PRESS

    0   0.1 0.12 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

    0.5

    1

    1.5

    −λ 

         µ   (       λ   )

    0.1 0.121.32

    1.34

    1.36

    1.38

    1.4

    1.42

    1.44

    1.461.48

    1.5

    Fig. 3. Front-tire friction coefficient for different values of  F z.

    -0.2

    0

    0.2

           λ   f

    0

    1000

    2000

       F  z  r   [   N   ]

    0

    10

    20

           θ

       [       °   ]

    4 4.5 5 5.5 6 6.5 70

    100

    200

    300

    time [s]

    4 4.5 5 5.5 6 6.5 7

    time [s]

    4 4.5 5 5.5 6 6.5 7

    time [s]

    4 4.5 5 5.5 6 6.5

    time [s]

      v   [   k  m   /   h   ]

    Rear-wheel loss of contact

    Constant slip target (-0.12)

    Fig. 4. Braking maneuver at constant target slip   ¯ l ¼ 0:12: the ‘‘stoppie’’ (from top: front-slip; rear vertical load; vehicle pitch angle; forward speed).

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657    647

  • 8/18/2019 optimal motorcycle braking

    5/14

    not consider the suspension dynamics; hence, it is valid for

    quasi-static conditions. The model is the following (sub-

    scripts ‘‘ f ’’ and ‘‘r’’ indicate front and rear parameters):

    m €x ¼ F x   C D   _x2;

    m€z ¼  F zr þ  F zf    mg þ C L   _x2;

    J y€y ¼  F zrb  F zf ð p  bÞ þ  F xh  C P  p _x2;

    8>>>: (2)where

      p, b, h   are the wheelbase, the distance between theprojection of the center-of-mass on the road and the

    rear-wheel contact point, and the height of the center-of-

    mass, respectively (see Fig. 2);

      {x, z, y} are the longitudinal, vertical and pitch coordi-nates of the vehicle center-of-mass, respectively;

      m, J y  are the vehicle mass and its pitch inertia (aroundthe center-of-mass), respectively;

      {C D, C L, C P } are the drag, lift and pitch aerodynamiccoefficients, respectively.

    Using the simple model (2), two useful quantities, as

    functions of the forward speed   _x, can be derived.

    The first is the deceleration   €xloss  which causes the loss of 

    contact of the rear wheel. It can be obtained by imposing€y ¼  0,   €z ¼  0,  F zr  ¼ 0. The result is the following:

    €xloss  ¼ mgð p  bÞ  C Lð p  bÞ _x

    2 þ C P  p _x2

    mh  þ C D   _x

    2. (3)

    The second is the deceleration   €xlock  which causes the wheel

    lock-up; it can be obtained by imposing   €y ¼  0,   €z ¼  0,

    F x  ¼ mmax   F zf , where   mmax ¼ mð0:12Þ. It is straightfor-ward to show that when braking only with the front brake,

    this deceleration is related to the forward speed   _x   as

    follows:

    €xlock  ¼  mmax

    ð p  mmaxhÞmðmgb  C L   _x

    2b  C P   _x2Þ þ C D   _x

    2. (4)

    The two variables (3)–(4) are plotted in   Fig. 5. The

    parameter values used in (3) and (4) have been extracted by

    the Bikesim model, in order to obtain consistent results.

    Note that in (3) and (4) the sign of    €x differs from (2), since

    the   €xloss  and   €xlock  are decelerations (they are positive when

    the vehicle speed decreases—see also  Fig. 5). The results,

    although somehow intuitive, are interesting: at high speed,

    thanks to the large aerodynamic forces, the wheel-lock

    limit is dominant; this means that all the friction capabilityof the front wheel can be used, and the slip-target policy is

    optimal. When the speed decreases, so do the aerodynamic

    forces, and the rear-wheel contact limit becomes active (in

    the simulated vehicle this limit is about 200 km/h, see also

    Fig. 4). That speed is the ‘‘critical-speed’’ from the rear-

    wheel contact point of view. Below that speed value, in

    order to minimize the stopping distance while maintaining

    the vehicle lateral stability, the controlled variable must be

    switched from the front wheel slip to the rear vertical load.

    Also notice that, at low speed,   €xloss   rapidly decreases, and

    the friction capability of the tires is significantly under-

    exploited.

    The optimal braking strategy in a motorbike hence is

    twofold: at high speed a slip-control must be used, the

    reference signal being the peak of the friction curve; below

    the critical speed this control policy must be disengaged

    and replaced with a controller of the rear-wheel vertical

    load; in practice, since the critical speed is hard to estimate,

    the switching variable is the rear-wheel load. The control

    target of the load controller must be ideally   ¯ F zr  ¼ 0þ; in

    practice, the zero-target has to be replaced with a small

    load (in the following, a load reference value of   ¯ F zr  ¼ 100

    has been used) to guarantee a minimum amount of 

    lateral contact force on the rear wheel. This control

    architecture is pictorially described in   Fig. 6. The resultsof this control policy are displayed in   Fig. 7. The two

    SISO controllers of  Fig. 6 are implemented with a standard

    PID architecture with anti-windup. Notice from   Fig. 7

    that the switch is seamless, since when the load-control

    PID is activated the state of its PI-part is not zero but it

    exactly equals the output of the slip controller at the

    switching time (this is implemented with the standard

    automatic-manual commutation algorithm for PID). In

    ARTICLE IN PRESS

    0 50 100 150 200 250 3000.8

    1

    1.8

    1.6

    1.4

    1.2

    2

    forward speed [km/h]

       l  o  n  g   i   t  u   d   i  n  a   l   d  e  c  e   l  e  r  a   t   i  o  n   [

      g   ]

    Rear-wheel contact limit  x loss

    Front-wheel-lock limit  x lock 

    Fig. 5. Deceleration limits at different speeds.

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657 648

  • 8/18/2019 optimal motorcycle braking

    6/14

    this way the switching guarantees a continuous behavior of 

    the control variable.

    From Fig. 7 notice that, when the rear-load controller is

    engaged, the front slip is progressively reduced, in order to

    avoid the loss of contact, while maximizing the braking

    performance. The braking maneuver of   Fig. 7   can be

    considered optimal when only the front brake is used. Also

    notice that—as expected—the switch of the target occurs at

    a forward speed slightly higher than the critical speed, since¯ F zr40.

    In the first part of this analysis the focus was on the

    general architecture of the control strategy. However, due

    to the complex dynamics of the vehicle and the tire, the

    very first part of the braking maneuver (say the first

    200 ms) is particularly critical. This issue is now analyzed in

    detail.

    The starting point of this analysis is the behavior of the

    automatic slip controller. This controller has been designed

    to provide the maximum achievable dynamic performance

    (maximum closed-loop bandwidth—see   Savaresi et al.,

    2007). The behavior of the control variable (front-braking

    torque) in the first 200 ms of the braking maneuver, when

    the target slip is set at   ¯ l ¼ 0:12, is displayed in   Fig. 8.Notice that the controller initially requires a ‘‘pulse’’ of 

    torque (about 10 ms long), due to the derivative part of the

    PID; then the torque is set to a small value and gradually

    increased again with a ramp-like behavior, until its steady-

    state value is reached. The ramp lasts about 150 ms. Notice,

    ARTICLE IN PRESS

    -

    -

    +

    +

    Vehicle

    dynamics

    Slip

    controller

    Load

    controller

    switch

    T  f 

     f 

    F  zr    F  zr 

    Fig. 6. Optimal control architecture for the front brake.

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.50

    100

    200

    300

    time [s]

      v   [   k  m   /   h   ]

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5-0.2

    -0.1

    0

    0.1

    0.2

    time [s]

           λ

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.50

    500

    1000

    1500

    2000

    time [s]

       F  z  r   [   N   ]

    front-slip target

    rear-load target

    slip-target control policy

    load-target control policy

    Fig. 7. Braking maneuver using the mixed slip-load control policy with only the front brake (from top: vehicle speed; front-wheel slip; rear-wheel vertical

    load).

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657    649

  • 8/18/2019 optimal motorcycle braking

    7/14

    however, that the control strategy implemented in  Fig. 8

    exceeds the actuator limit (saturation), and, in practice, the

    initial pulse is cut-off.

    Starting from this ideal behavior, a sensitivity analysis

    on the effects of three different control strategies in this

    initial phase of the braking maneuver has been carried out.

    The three initial control strategies are (see the top sub-plotof  Fig. 9):

     Control strategy 1. This control strategy simply mimicsthe behavior of the closed-loop slip controller: an initial

    short (10 ms) pulse, using the maximum available

    torque, is followed by a 140 ms ramp which starts from

    a value of about 40% of the final steady-state value.

      Control strategy 2. This control strategy mimics thebehavior of a ‘‘real-driver’’ (see   Fig. 1): the braking

    torque is increased from 0% to 100% of the steady-state

    value, with a 150 ms ramp (notice that a very fast-reactingdriver has been assumed; in the measurement displayed in

    Fig. 1 the driver takes 250 ms to make the ramp).

      Control strategy 3. The idea of this control strategy isto immediately set the braking torque to the right

    ARTICLE IN PRESS

    3.95 4 4.05 4.1 4.15 4.20

    1000

    2000

    3000

    4000

    5000

    6000

    7000

    8000

    9000

    time [s]

       T   [

       N  m   ]

    Front Braking Torque

    Actuator limit(saturation)

    Initial pulse(withoutsaturation)

    Torque ramp Steady-state torque

    Fig. 8. Initial behavior of the control variable (front braking torque) using a large-bandwidth slip controller.

    4 4.05 4.1 4.15 4.2 4.250

    1000

    2000

    3000

    time [s]

    4 4.05 4.1 4.15 4.2 4.25

    time [s]

    4 4.05 4.1 4.15 4.2 4.25

    time [s]

    4 4.05 4.1 4.15 4.2 4.25

    time [s]

       T   f   [   N  m   ]

    -0.2

    -0.1

    0

           λ   f

    0

    1000

    2000

       F  z  r   [   N

       ]

    285

    290

    295

    300

      v   [   k  m   /   h   ]

    Control strategy 1

    Control strategy 2

    Control strategy 3

    Performance difference: about 30ms

    Fig. 9. Sensitivity analysis to initial transient control strategy (from top: front braking torque; front-wheel slip; rear-wheel vertical load; vehicle speed).

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657 650

  • 8/18/2019 optimal motorcycle braking

    8/14

    steady-state value; in practice this would cause wheel

    lock, since the load cannot be transferred to the front-

    wheel instantaneously, due to the suspension dynamics;

    hence, in practice, the braking torque is set to about

    80% of the steady-state value, followed by a 150 ms

    ramp which closes the 20% gap.

    The results of these three different control policies are

    condensed in Fig. 9, where the longitudinal slip, rear load,

    and forward speed of the vehicle are displayed. The

    following remarks are due:

      Even if the control strategies 1 and 3 show a remarkablydifferent behavior in the longitudinal slip, their overall

    performances are very similar, in terms of vehicle

    deceleration. This can be explained with the fact that

    in the very initial part of the braking maneuver

    (20–30 ms) the load has not yet been fully transferred

    to the front wheel; thus, during this phase, the sensitivity

    to different values of wheel slip is comparatively small.

      The control strategy 2 shows significantly lowerperformance than 1 and 3. This is clearly due to the

    fact that the friction capability of the tire is under-

    exploited, since the magnitude of the wheel slip is

    increased too slowly.

    It is interesting to observe that the control strategy 2 is the

    worst from the performance point of view, but it is the

    most natural and easiest for a ‘‘human’’ controller (see

    Fig. 1). On the other hand, control strategies 1 and 3

    provide higher performance, but they can hardly be

    managed by a human driver. Note that the difference of performance in the 300-290 km/h deceleration is about

    30ms; it is non-negligible, considering that all this

    difference is condensed in the very first part of the braking

    maneuver. Hence, this implicitly confirms the potentially

    beneficial effect of an electronic brake assistant.

    3. Optimal front–rear braking

    In racing motorcycles the role and the importance of rear

    braking is very debated as, in practice, during a hard

    braking on a straight line, most of the drivers make no use

    at all of the rear brake. This brake is typically used for

    better managing the attitude and the stability of the

    motorbike during a bend. The reason is threefold:

     Due to the (almost) total load transfer to the front tire atmid-low speed in a hard-braking maneuver, the braking

    capability of the rear tire is comparatively small.

     The simultaneous optimal management of the front andrear brake is a very hard task even for a professional

    driver.

     If the engine is engaged during braking, the load torqueof the engine (especially on high-performance high-

    compression 4-stroke engines) in many cases is high

    enough to lock the rear wheel; this phenomenon can be

    alleviated using mechanical devices called ‘‘anti-hop-

    ping’’ clutches, which can be considered as a raw anti-

    lock braking system acting on the engine-induced

    braking torque.

    In order to simplify the analysis and to give a deep

    understanding of the role of the rear brake, it is assumedthat the engine is disengaged during braking, and that all

    the rear-braking torque is provided by the rear brake.

    As a starting point, it is interesting to analyze the

    behavior of the rear-wheel slip when the front-brake mixed

    slip-load control strategy illustrated in   Figs. 6 and 7   (see

    the previous section) is used. The rear and front slip are

    displayed in Fig. 10.

    From Fig. 10 it is interesting to point out that, at the rear

    wheel, the slip is  positive, namely the rear wheel provides a

    traction   torque, not a braking torque. This fact can be

    somewhat surprising at a first glance, but it can be easily

    explained by observing that the rear wheel has a kinetic

    energy given by   12J rð ¯ orÞ2, where   ¯ or   is the rear wheel

    rotational speed at the beginning of the braking maneuver,

    and   J r   the inertia of the rear wheel. If the engine is

    disengaged during braking and only the front brake is used,

    this energy is not dissipated but it is transformed into a

    traction torque.

    When the rear brake is used, the control architecture of 

    Fig. 6   must be changed into the control architecture of 

    Fig. 11. The mixed slip-load strategy at the front brake is

    kept unchanged; the rear wheel is simply managed by

    regulating its slip  lr  to the peak value of the rear friction

    curve.

    The results obtained with this control architecture aredisplayed in Fig. 12. These results are compared with those

    obtained using the front brake only. The major difference

    is, obviously, in the slip of the rear wheel. As already

    noticed, this slip is positive (namely it provides   traction

    torque) if no rear brake is used; if the rear brake is used, the

    rear slip is regulated to the same target   ð¯ l ¼ 0:12Þ   usedfor the front slip. Notice that the regulation of the rear slip

    is a difficult control task, due to the very small load on the

    rear wheel which causes a significant cross-disturbance of 

    the front-brake controller onto the rear-slip dynamics. This

    non-perfect tracking of the rear wheel slip, however, has

    little effect on the overall performance.

    This section is concluded by observing that in the long

    300-80 km/h braking maneuver, the difference of perfor-

    mance is very large, about 300 ms. This difference is largely

    due to the fact that the ‘‘traction’’ torque has been replaced

    by a ‘‘braking’’ torque at the rear wheel Although the

    braking torque at the rear wheel is rather small compared

    to the front-wheel one, on a strong braking maneuver the

    effect of the rear brake can be clearly appreciated. This is

    true, in particular, if the front-brake controller is able to

    maintain a small but non-zero load on the rear tire.

    For the sake of completeness it is worth mentioning that,

    in practice, this performance improvement is significantly

    reduced if the engine is engaged and the vehicle is equipped

    ARTICLE IN PRESS

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657    651

  • 8/18/2019 optimal motorcycle braking

    9/14

    ARTICLE IN PRESS

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5

    -0.2

    -0.1

    0

    0.1

    0.2

    time [s]

           λ

    Front λ

    Rear λ

    Target front λ

    slip-target control policy

    load-target control policy

    Fig. 10. Front- and rear-wheel slip using a front-brake mixed slip-load control policy.

    -+

    -+

    Vehicle

    dynamics

     f 

     f Slip

    controller

    Load

    controller

    switch

    T  f 

      -

    + Slipcontroller

    F  zr    F  zr 

    Fig. 11. Optimal control architecture of the front and rear brakes.

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5

    100

    150

    200

    250

    300

      v   [   k  m   /   h   ]

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5

    4 4.5 5 5.5 6 6.5 7 7.5 8 8.5 9 9.5

    -0.2

    0

    0.2

           λ   f

    -0.2

    0

    0.2

           λ  r

    0

    1000

    2000

    time [s]

       F  z  r   [   N   ]

    Front Brake Only

    Both Brakes

    Performance difference: about 300ms

     = -0.12

     = -0.12

    Fig. 12. Brake maneuver using the complete mixed slip-load control policy using the front and the rear brake (from top: longitudinal speed; front slip; rear

    slip; rear vertical load).

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657 652

  • 8/18/2019 optimal motorcycle braking

    10/14

    with anti-hopping clutch. From this point of view, the anti-

    hopping clutch can be interpreted as a very raw automatic

    rear slip controller. Obviously enough, a genuine electronic

    slip controller (acting on an electronically-controlled clutch

    or brake) can guarantee much better and more consistent

    slip-tracking performances.

    This analysis demonstrates the potential benefits of anautomatic control strategy for the rear brake: the braking

    performance is maximized while guaranteeing the road

    contact of the rear wheel. A somehow futuristic but very

    attractive way of achieving such high performance is the

    by-wire electro-mechanical brake (see e.g.,  Savaresi et al.,

    2007).

    4. Sensitivity to front-suspension damping

    In this section the effect of the damping ratio of the front

    suspension (fork) during a braking maneuver is discussed

    (see also Cossalter et al., 2000). As a matter of fact, amongthe many tunable parameters of a motorbike, the damping

    ratio of the fork is one of the most critical, since the vehicle

    dynamics are very sensitive to this parameter, and there is a

    strong trade-off in its tuning (see below). On this

    parameter, drivers and race engineers are always struggling

    in the search for the best compromise among different

    conflicting requirements.

    The first part of the analysis has been developed in the

    frequency domain; more specifically, two transfer functions

    have been experimentally estimated using the simulator:

     the transfer function from the front braking torque ( T  f  )to the front wheel longitudinal slip (l f  ); this transferfunction is useful in the design of slip controllers;

     the transfer function from the front braking torque ( T  f  )to the front wheel vertical contact force   ðF z f  Þ; this

    transfer function is useful to investigate the minimiza-

    tion of vertical-load variations, which is a classical

    prerequisite to achieve high (and constant) longitudinal

    and lateral contact forces.

    Since the braking maneuver is a non-equilibrium condition

    (due to vehicle deceleration), the computation of such

    transfer functions cannot be managed with traditionallocal-linearization tools, which typically require a well-

    defined steady-state condition; thus, an ad-hoc small-

    perturbation approach has been used (see e.g.,   Guarda-

    bassi & Savaresi, 2001; Savaresi &  Spelta, 2007).

    Fig. 13  shows a single step of this procedure: starting

    from a vehicle in straight running with a speed of 200 km/h,

    a constant front braking torque is applied; this constant

    torque generates (after a short transient) a steady-state

    front-wheel slip. Then a small sinusoidal perturbation is

    superimposed to the constant torque (a 2 Hz perturbation

    is depicted in Fig. 13); the corresponding slip is affected by

    a sinusoidal oscillation at the same frequency; by compar-ing the amplitude and phase relationships between the two

    sinusoidal oscillations, a single point of the estimated

    frequency response can be computed. This procedure has

    been repeated for a large number of frequencies (from 0.5

    to 15 Hz, with a frequency resolution of 0.25Hz).

    The magnitudes of the estimated transfer functions are

    displayed in Fig. 14, for three different values of damping

    (low: 1000 Ns/m; medium: 3000 Ns/m; high: 5000 Ns/m).

    The following remarks can be made.

     In all cases, a strong trade-off is clearly visible: at low-

    damping all the transfer functions are characterized by alow-frequency (2.5 Hz) poorly-damped resonance; by

    increasing the damping ratio this resonance phenomen-

    on first disappears (for medium damping) and then

    reappears as a poorly-damped resonance located at

    about 7 Hz.

    ARTICLE IN PRESS

    1 2 3 4 5 6 7 8 9 10

    0

    100

    200

    300

    400

       T   f   [   N  m   ]

    1 2 3 4 5 6 7 8 9 10-0.04

    -0.02

    0

           λ   f

    1 2 3 4 5 6 7 8 9 10

    100

    150

    200

    time [s]

      v   [   k  m   /   h   ]

    2Hz sinusoidal perturbation on Tf 

    2Hz sinusoidal response on λ f 

    Fig. 13. Example of single-tone perturbation for frequency–response estimation.

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657    653

  • 8/18/2019 optimal motorcycle braking

    11/14

      The resonance in the slip response is critical for thecontrol of the wheel slip (both from a human-driver

    perspective and from an automatic-controller perspec-

    tive). The resonance in the load-response to the braking

    torque is critical for the behavior of the contact forces.

      In both cases, for low (below 5 Hz) and high (beyond

    9 Hz) frequencies, the high-damping configuration is thebest. The low-damping configuration outperforms it in

    the mid-range frequency range (5–9 Hz) only. This

    trade-off is well known for the vertical movement of 

    the sprung and unsprung masses; it is interesting to see

    that this behavior also holds for the slip dynamics of 

    the tire.

    The frequency-domain interpretation has been comple-

    mented with a time-domain analysis. To this end, in Fig. 15

    the time responses of vertical contact force (F z), long-

    itudinal contact force (F x) and fork elongation (Dx), to a

    step in the front braking torque, are displayed for the three

    different levels of damping of the front suspension. All

    these responses refer to the front axle. By analyzing thesestep responses, additional insight in the influence of the

    damping ratio during braking can be gained. Some

    remarks are due.

     The first remark refers to the initial step response of thevertical contact force, in case of low damping. Notice a

    ARTICLE IN PRESS

    100

    100

    101

    101

    -90

    -85

    -80

    -75

    -70

      g  a

       i  n   [   d   B   ]

      g  a   i  n   [   d   B   ]

    Estimated transfer function from front-braking-torque (Tf ) to front-tire slip (λ f )

    -10

    0

    10

    20

    30

    Estimated transfer function from front-braking-torque (Tf ) to vertical load at the front-tire contact point (Fzf )

    Best damping: high

    Best damping: high

    Best damping: high

    Best damping: high

    Best damping: low

    Best damping: low

    Frequency [Hz]

    Low damping

    Medium damping

    High damping

    Fig. 14. Frequency-domain analysis of the sensitivity of braking torque responses to fork damping.

    1.5 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3

    1500

    2000

    2500

       F  z   [   N   ]

    1.5 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3

    0

    500

    1000

    1500

    2000

    2500

       F  x   [   N   ]

    1.5 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3

    0

    20

    40

    60

    80

    time [s]

          ∆  x   [  m  m   ]

    Low dampingMedium dampingHigh damping

    Fast compression rate

    Slow compression rate

    Fig. 15. Responses (from top: vertical and longitudinal contact forces; fork elongation) to a step on the front braking torque.

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657 654

  • 8/18/2019 optimal motorcycle braking

    12/14

    seemingly strange behavior: when a braking torque step

    is applied, the contact force initially decreases; then it

    increases and reaches its steady-state value. This

    behavior is not intuitive, since a load transfer from the

    rear to the front wheel is expected, whatever the

    damping ratio of the fork is. This unusual phenomenon

    is due to the fact that the angle with respect to the roadsurface of the front suspension of a motorbike

    significantly differs from 901. The angle between the

    fork and a vertical line—called caster angle—for a sport

    motorbike is usually about 251 (26.11 in our case). When

    a step on the braking torque is applied, a non-zero

    longitudinal braking force is rapidly generated; due to

    the non-zero caster angle, this force is not perpendicular

    to the front suspension; hence, the component of this

    force projected onto the fork axis tends to compress the

    suspension, unloading the tire (see   Fig. 16). This

    phenomenon is particularly visible if the damping ratio

    is low. Obviously, this phenomenon is rapidly compen-

    sated by the load transfer from the rear to the front

    wheel. This remark shows that, at the very beginning of 

    the braking maneuver, the best setting for the damping

    ratio is ‘‘high’’.

      From the stopping-time point of view, the mostrepresentative variable in a braking maneuver is the

    longitudinal contact force  F x, since it is the actual force

    which decelerates the vehicle. Hence, this force must be

    maximized at all times. By comparing the  F x   responses

    at low and high damping, it is interesting to observe

    that, in both cases, a resonance phenomenon occurs. In

    the case of low damping this phenomenon occurs at low

    frequency; in the case of high damping, at a higherfrequency. This obviously confirms the frequency

    domain analysis. It is interesting to point out that the

    overall braking performance is indeed very similar with

    both damping values: a careful inspection of the two

    responses shows that the overall balance of the long-

    itudinal force is approximately even, whatever the

    damping ratio value is. Notice that this substantial

    independence of the braking performance from the

    damping ratio no longer holds if a dynamic adjustment

    of the damping can be done (semi-active damping

    control—e.g.,   Savaresi, Bittanti,   &   Montiglio, 2005;

    Savaresi   &   Spelta, 2007). It is easy to see that a

    ‘‘high–medium–high’’ (or even ‘‘high–low–high’’) damp-ing switch during the first 300 ms of the braking

    maneuver would maximize the overall longitudinal

    forces. This brake-oriented semi-active control strategy

    (which is out of the scope of the present work) is

    currently under development.

    Another important characteristic to be carefully consid-

    ered during braking is the behavior of the fork compression,

    since this variable is directly ‘‘sensed’’ by the driver. By

    inspecting its step response for different damping ratios, it is

    clear that, with the low damping setting, the fork rapidly

    reaches its steady-state value. However, this fast transient is

    paid for overshoot and under-damped oscillations. On the

    other hand, the high damping setting guarantees no

    overshoot and no oscillations. In this case, the price to be

    paid is a slower response: the fork takes almost 500 ms to

    reach its steady-state condition. In a short braking

    maneuver followed by a bend, this slow transient is

    considered as a drawback, since the driver typically wants

    to ‘‘feel’’ a steady-state condition before engaging a turn.

    Also in this case, semi-active damping can help to alleviate

    the trade-off: notice that the ‘‘high–medium–high’’ switch-

    ing policy above described can also speed up the transient

    of the high damping setting, while maintaining the

    attractive properties of no overshoot, no oscillations andno initial unload of the front wheel.

    5. Conclusions and future work

    In this work the problem of high-performance braking

    has been studied. The focus is on optimizing the braking

    performance, in terms of stopping time. All the main

    aspects of the problem have been considered and discussed:

    the effect of aerodynamics on the loss of contact of the rear

    wheel; the wheel-lock problem; the best torque modulation

    strategy at the beginning of the braking maneuver; and the

    role of the rear brake. Moreover, the sensitivity of the

    braking performance to fork damping has been studied.

    The results of this analysis provide some useful insights

    on the problem of optimal braking. Moreover, this analysis

    shows the potential benefits of electronically assisted

    brakes: these benefits could be further exploited by joint

    design of braking control and semi-active suspension

    damping control.

    Acknowledgments

    This work has been supported by  MIUR PRIN project

    ‘‘identification and adaptive control of industrial systems’’.

    Thanks are due to Mario Santucci, Andrea Strassera and

    ARTICLE IN PRESS

    F  x  f 

    Casterangle

    Fig. 16. A pictorial representation of the fork compression due to the

    braking force and the caster angle.

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657    655

  • 8/18/2019 optimal motorcycle braking

    13/14

    Onorino di Tanna of  Piaggio Group, Carlo Cantoni and

    Roberto Lavezzi of   Brembo, for enlightening discussions

    on the topic of braking control.

    Appendix A. Motorcycle geometric and mechanical

    properties

    Wheelbase 1.370 m

    Center of mass height (vehicle with rider) 0.621 m

    Center of mass distance from front wheel 0.560 m

    Caster angle 26.11

    Rear arm length 0.250 m

    Mass of vehicle with rider 274.2 kg

    Front wheel mass 12.7 kg

    Rear wheel mass 14.7 kg

    Frame pitch inertia 22 kgm2

    Front wheel moment of inertia 0.484 k gm2

    Rear wheel moment of inertia 0.638 kgm2

    Frontal cross section area 0.6 m2

    Drag coefficient 0.52Lift coefficient 0.085

    Pitch coefficient 0.205

    Rear suspension properties

    Spring stiffness 40,000 N/m

    Spring pre-load 2080 N

    Spring travel 0.2 m

    Damping coefficient 10,000 Ns/m

    End stroke stiffness 600,000 N/m

    Swing arm mass 8 kg

    Front suspension properties

    Spring stiffness 40,000 N/m

    Spring pre-load 933 N

    Spring travel 0.12 m

    Damping coefficient 2500 Ns/m

    End stroke stiffness 100,000 N/m

    Front fork mass 7.25 kg

    Tire properties

    Radial stiffness of front tire

    130,000 N/m

    Radial stiffness of 

    rear tire

    141,000 N/m

    Paceika parameters

    of front tire

    120/70, (Sharp, Evangelou,

    & Limebeer, 2004)

    Paceika parameters

    of rear tire

    180/55, (Sharp, Evangelou,

    & Limebeer, 2004)

    References

    Buckholtz, K.R. (2002). Reference input wheel slip tracking using sliding

    mode control. SAE technical paper 2002-01-0301.

    Cossalter, V. (2002).   Motorcycle dynamics. USA: Race Dynamics,

    Milwaukee.

    Cossalter, V., Doria, A., & Lot, R. (1999). Steady turning of two-wheeled

    vehicles.  Vehicle System Dynamics,  31, 157–181.

    Cossalter, V., & Lot, R. (2002). A motorcycle multi-body model for real

    time simulations based on the natural coordinates approach.   Vehicle

    System Dynamics: International Journal of Vehicle Mechanics and 

    Mobility,  37 , 423–447.Cossalter, V., Lot, R., & Maggio, F. (2002). The Influence of tire

    properties on the stability of a motorcycle in straight running and in

    curve. In  SAE automotive dynamics and stability conference (ADSC),

    Detroit, USA.

    Cossalter, V., Lot, R., & Maggio, F. (2004). On the stability of motorcycle

    during braking. In   SAE small engine technology conference and 

    exhibition, Graz, Austria, SAE paper number: 2004-32-0018/

    20044305.

    Cossalter, V., Doria, A., & Lot, R. (2000). Optimum suspension design for

    motorcycle braking. Vehicle System Dynamics: International Journal of 

    Vehicle Mechanics and Mobility,  34, 175–198.

    Donida, F., Ferretti, G., Savaresi, S. M., Schiavo, F., & Tanelli, M.

    (2006). Motorcycle dynamics library in Modelica, Modelica Conference.

    Frezza, R., & Beghi, A. (2005). Simulating a motorcycle driver. New trends

    in nonlinear dynamics and control, and their applications   (Vol. 295,pp. 175–187). ISBN: 3-540-40474-0. Lecture notes in control and

    information sciences. Berlin: Springer.

    Frezza, R., Beghi, A., & Saccon, A. (2004). A model predictive control for

    path following with motorcycles: Application to the development of 

    the pilot model for virtual prototyping. In   43rd IEEE conference on

    decision and control  (pp. 767–772).

    Guardabassi, G. O., & Savaresi, S. M. (2001). Approximate lineariza-

    tion via feedback—An overview.   Survey paper on Automatica,   27 ,

    1–15.

    Hauser, J., & Saccon, A. (2006). Motorcycle modeling for high-

    performance maneuvering.   IEEE Control Systems Magazine,   26 (5),

    89–105.

    Hauser, J., Saccon, A., & Frezza, R. (2004). Achievable motorcycle

    trajectories. In   43rd IEEE conference on decision and control 

    (pp. 3944–3949).Johansen, T. A., Petersen, J., Kalkkuhl, J., & Lu ¨ demann, J. (2003). Gain-

    scheduled wheel slip control in automotive brake systems.   IEEE 

    Transactions on Control Systems Technology,  11(6), 799–811.

    Limebeer, D. J. N., Sharp, R. S., & Evangelou, S. (2001). The stability of 

    motorcycles under acceleration and braking.  Institution of Mechanical 

    Engineering, Part C, Journal of Mechanical Engineering Science,  215,

    1095–1109.

    Lin, C.-F., Chyuan-Yow, T., & Tsai-Wen, T. (2006). A hardware-in-the-

    loop dynamics simulator for motorcycle rapid controller prototyping.

    Control Engineering Practice,  14(12), 1467–1476.

    Pacejka, H. B. (2002).  Tyre and vehicle dynamics. Oxford: Buttherworth

    Heinemann.

    Savaresi, S. M., Bittanti, S., & Montiglio, M. (2005). Identification of 

    semi-physical and black-box non-linear models: The case of MR-

    dampers for vehicles control.   Automatica,  41, 113–117.Savaresi, S. M., & Spelta, C. (2007). Mixed Sky-Hook and ADD:

    Approaching the filtering limits of a semi-active suspension. ASME

    transactions:  Journal of Dynamic Systems, Measurement and Control ,

     29(4), 382–392.

    Savaresi, S. M., Tanelli, M., & Cantoni, C. (2007). Mixed slip-deceleration

    control in automotive braking systems.   ASME Transactions:

    Journal of Dynamic Systems, Measurement and Control ,   129(1),

    20–31.

    Savaresi, S. M., Tanelli, M., Langthaler, P., & Del Re, L. (2006).

    Estimation of X–Z tire contact force via embedded accelerometers.

    Tire Technology International , 74–78.

    Sayers, M. W. (1999). Vehicle models for RTS applications.   Vehicle

    System Dynamics,  32, 421–438.

    Sharp, R. S. (1971). The stability and control of motorcycles.  Journal of 

    Mechanical Engineering Science,  13, 316–329.

    ARTICLE IN PRESS

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657 656

  • 8/18/2019 optimal motorcycle braking

    14/14

    Sharp, R. S. (2001). Stability, control and steering responses of 

    motorcycles.  Vehicle System Dynamics,  35, 291–318.

    Sharp, R. S., Evangelou, S., & Limebeer, D. J. N. (2004). Advances in the

    modelling of motorcycle dynamics.  Multibody System Dynamics,   12,

    251–283.

    Sharp, R. S., Evangelou, S., & Limebeer, D. J. N. (2005). Multibody

    aspects of motorcycle modelling with special reference to Autosim.

    In J. A. C. Ambrosio (Ed.),   Advances in computational multibody

    systems (pp. 45–68). Dordrecht: Springer.

    Sharp, R. S., & Limebeer, D. J. N. (2001). A motorcycle model

    for stability and control analysis.   Multibody System Dynamics,   6 ,

    123–142.

    Solyom, S., Rantzer, A., & Lu ¨ demann, J. (2004). Synthesis of a model-

    based tire slip controller.  Vehicle System Dynamics,  41(6), 477–511.

    ARTICLE IN PRESS

    M. Corno et al. / Control Engineering Practice 16 (2008) 644–657    657