23.11.2012 Views

04 Steering And Turning Vehicles - Department of Mechanical ...

04 Steering And Turning Vehicles - Department of Mechanical ...

04 Steering And Turning Vehicles - Department of Mechanical ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Steering</strong> and <strong>Turning</strong> <strong>Vehicles</strong><br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Pr<strong>of</strong>. R.G. Longoria<br />

Spring 2012<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


• <strong>Steering</strong> mechanisms<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Overview<br />

• Differentially-steered vehicle – kinematic<br />

• Ackermann steered vehicle – kinematic<br />

• Simulation examples<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Classical steering mechanisms<br />

δ =<br />

H<br />

5 th wheel<br />

steering<br />

'hand wheel' angle<br />

‘turntable steering’<br />

•Likely developed by the Romans, and preceded only by a 2<br />

wheel cart.<br />

•Consumes space<br />

•Poor performance – unstable<br />

•Longitudinal disturbance forces have large moment arms<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

•Articulated-vehicle steering<br />

•Tractors, heavy industrial<br />

vehicles<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Common <strong>Steering</strong> Mechanisms<br />

Differential steer Synchro-drive Tricycle<br />

What is minimum<br />

# <strong>of</strong> actuators?<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

…and some systems also employ Ackermann-type.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Kinematic model <strong>of</strong> single-axle turning vehicle<br />

Y<br />

Recall the simple 2D turning vehicle with kinematic state quantified by,<br />

y<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

x<br />

ψ<br />

X<br />

X<br />

[ ] †<br />

qI<br />

= X Y ψ<br />

The velocities in the local (body-fixed) reference frame<br />

are transformed into a global frame by the rotation<br />

matrix,<br />

⎡ cosψ sinψ 0⎤<br />

R(<br />

ψ ) =<br />

⎢<br />

−sinψ<br />

cosψ 0<br />

⎥<br />

⎢ ⎥<br />

⎢⎣ 0 0 1⎥⎦<br />

and the velocities are then,<br />

Inverting, we arrive at the velocities in the global reference frame,<br />

qɺ = R( ψ ) ⋅qɺ<br />

I<br />

qɺ = Ψ( ψ ) ⋅qɺ<br />

where,<br />

⎡cosψ −sinψ<br />

0⎤<br />

Ψ(<br />

ψ ) =<br />

⎢<br />

sinψ cosψ 0<br />

⎥<br />

⎢ ⎥<br />

⎢⎣ 0 0 1⎥⎦<br />

Let’s apply these relations to the case <strong>of</strong> a single-axis vehicle that has two wheels<br />

differentially driven with controlled speed.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin<br />

I


Y<br />

Recall: position and velocity in inertial frame<br />

We defined the vehicle’s kinematic state in the inertial frame by,<br />

l2<br />

y<br />

B<br />

l1<br />

x<br />

ψ<br />

So, for our simple (single-axle) vehicle,<br />

the velocities in the inertial frame in<br />

terms <strong>of</strong> the wheel velocities are,<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

X<br />

X<br />

Velocities in the local (body-fixed)<br />

reference frame are transformed<br />

into the inertial frame by the<br />

rotation matrix,<br />

or, specifically, qɺ = R( ψ ) ⋅qɺ<br />

I<br />

q<br />

I<br />

⎡ X ⎤<br />

=<br />

⎢<br />

Y<br />

⎥<br />

⎢ ⎥<br />

⎢⎣ ψ ⎥⎦<br />

⎡ cosψ sinψ 0⎤<br />

R(<br />

ψ ) =<br />

⎢<br />

−sinψ<br />

cosψ 0<br />

⎥<br />

⎢ ⎥<br />

⎢⎣ 0 0 1⎥⎦<br />

Inverting, we arrive at the velocities in the global reference<br />

frame,<br />

⎡ Xɺ ⎤ ⎡U ⎤ ⎡cosψ ⎢ ⎥<br />

qɺ I = Yɺ ⎢<br />

V<br />

⎥<br />

( ψ )<br />

⎢<br />

⎢ ⎥ =<br />

⎢ ⎥<br />

= Ψ ⋅ qɺ<br />

=<br />

⎢<br />

sinψ ⎢ψ ⎥<br />

⎣<br />

ɺ<br />

⎦<br />

⎢⎣ Ω⎥⎦<br />

⎢⎣ 0<br />

−sinψ<br />

cosψ 0<br />

0⎤<br />

⎡ vx<br />

⎤<br />

0<br />

⎥ ⎢<br />

v<br />

⎥<br />

⎥ ⎢ y ⎥<br />

1⎥⎦<br />

⎢⎣ ω ⎥ z ⎦<br />

qɺ<br />

I<br />

⎡ Rw l2Rw ⎤<br />

⎢ ( ω1 + ω2)cos ψ − ( ω1 −ω2<br />

)sinψ<br />

2<br />

B<br />

⎥<br />

⎡Xɺ ⎤ ⎢ ⎥<br />

⎢ ⎥ Rw l2Rw = Yɺ<br />

⎢ ( ω1 ω2)sin ψ ( ω1 ω2)cos ψ ⎥<br />

⎢ ⎥ = + + −<br />

⎢ 2<br />

B<br />

⎥<br />

⎢ψ ⎥<br />

⎣<br />

ɺ<br />

⎦ ⎢ ⎥<br />

⎢<br />

Rw<br />

( ω1 −ω2<br />

)<br />

⎥<br />

⎢⎣ B<br />

⎥⎦<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Example: Differentially-driven basic vehicle<br />

For a kinematic model <strong>of</strong> a differentially-driven vehicle, we assume there is no slip, and<br />

that the wheels have controllable speeds, ω 1 and ω 2 . If the CG is on the rear axle,<br />

the velocities in the global reference frame are,<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Y<br />

y<br />

X<br />

x<br />

ψ<br />

B = track width<br />

qɺ<br />

I<br />

l = L<br />

l<br />

1<br />

2<br />

=<br />

0<br />

⎡ Rw<br />

⎤<br />

⎢ ( ω1 + ω2) cosψ<br />

2<br />

⎥<br />

⎡Xɺ ⎤ ⎢ ⎥<br />

⎢ ⎥ Rw<br />

= Yɺ<br />

⎢ ( ω1 ω2)sin ψ ⎥<br />

⎢ ⎥ = +<br />

⎢ 2<br />

⎥<br />

⎢ψ ⎥<br />

⎣<br />

ɺ<br />

⎦ ⎢ ⎥<br />

⎢<br />

Rw<br />

( ω1 −ω2<br />

) ⎥<br />

⎢⎣ B<br />

⎥⎦<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Simulation <strong>of</strong> the differentially-steered single-axle vehicle trajectory<br />

A simple code in Matlab to compute and plot out the vehicle trajectory is given below.<br />

% Differentially-steered kinematic vehicle model<br />

% Requires right (#1) and left (#2) wheel velocities, omegaw1 and omegaw2,<br />

% as controlled inputs for single axle, to be passed as global parameters<br />

% Wheel radius, R_w, and axle track width, B, are also required<br />

% Updated 2/20/12 RGL<br />

function Xidot = DS_vehicle(t,Xi)<br />

global R_w B omegaw1 omegaw2<br />

X = Xi(1); Y = Xi(2); psi = Xi(3);<br />

% NOTE: these are global coordinates<br />

% These equations assume CG on single axle<br />

Xdot = 0.5*cos(psi)*R_w*(omegaw1+omegaw2);<br />

Ydot = 0.5*sin(psi)*R_w*(omegaw1+omegaw2);<br />

psidot = R_w*(omegaw1-omegaw2)/B;<br />

Xidot=[Xdot;Ydot;psidot];<br />

% test_DS_vehicle.m<br />

clear all<br />

global R_w B omegaw1 omegaw2<br />

% Rw = wheel radius, B = track width<br />

% omegaw1 = right wheel speed<br />

R_w = 0.05; B = 0.18;<br />

omegaw1 = 4; omegaw2 = 2;<br />

Xi0=[0,0,0];<br />

[t,Xi] = ode45(@DS_vehicle,[0 10],Xi0);<br />

N = length(t);<br />

figure(1)<br />

plot(Xi(:,1),Xi(:,2)), axis([-1.0 1.0 -0.5 1.5]), axis('square')<br />

xlabel('X'), ylabel('Y')<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Vehicle trajectory in XY<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Simulation <strong>of</strong> differentially-steered vehicle with 2D animation<br />

A code in Matlab to plot out the vehicle trajectory including a simple graphing animation <strong>of</strong><br />

the vehicle body/orientation is provided on the course log. This provides visual feedback on<br />

the model results.<br />

The key elements <strong>of</strong> this code are:<br />

Y<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.5<br />

-1 -0.5 0<br />

X<br />

0.5 1<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

1. Specify and plot initial location<br />

and orientation <strong>of</strong> the vehicle CG.<br />

2. Initiate some ‘handle graphics’<br />

functions for defining the ‘body’.<br />

3. Perform a fixed wheel speed<br />

simulation loop to find state, q.<br />

4. The state <strong>of</strong> the robot is used to<br />

define the position and orientation<br />

<strong>of</strong> the vehicle over time.<br />

5. A simple routine is used to<br />

animate 2D motion <strong>of</strong> the vehicle<br />

by progressive plotting <strong>of</strong> the<br />

body/wheel positions.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Lankensperger/‘Ackermann’-type steering<br />

‘knuckle’<br />

1. <strong>Steering</strong> arm<br />

2. Drag link<br />

3. Idler arm<br />

4. Tie rod/rack<br />

5. <strong>Steering</strong> wheel<br />

6. <strong>Steering</strong> shaft<br />

7. <strong>Steering</strong> box<br />

8. Pitman arm<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Rigid axle with<br />

kingpin<br />

Divided track rods<br />

for independent<br />

suspension.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


<strong>Turning</strong> at low (Ackermann) speed<br />

• What is low-speed?<br />

– Negligible centrifugal forces<br />

– Tires need not develop lateral<br />

forces<br />

• Pure rolling, no lateral sliding<br />

(minimum tire scrub).<br />

• For proper geometry in the turn,<br />

the steer angles, δ, are given by:<br />

L L<br />

δo ≅ < δi<br />

≅<br />

R + B R − B<br />

2 2<br />

• The average value (small angles)<br />

is the Ackerman angle,<br />

δ =<br />

Ackermann<br />

L<br />

R<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

L<br />

δ o<br />

B<br />

δ i<br />

R<br />

Simple relationship between<br />

heading and steering wheel angle.<br />

Ref. Wong, Ch. 5<br />

Turn Center<br />

“Ackermann steering geometry”<br />

cotδ o − cotδ<br />

i = B<br />

L<br />

δ = steer angle <strong>of</strong> outside wheel<br />

o<br />

δ = steer angle <strong>of</strong> inside wheel<br />

i<br />

B = track<br />

L = wheelbase<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Wheel slip angle definition<br />

A positive slip angle is said to give a<br />

negative force on the wheel (to the left).<br />

Gillespie (1992)<br />

*To align the wheel with heading direction<br />

(vector), you would have to go –α.<br />

The lateral force should simply always<br />

oppose the motion – the lateral force is a<br />

dissipative force!<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Fy = Cα ⋅α<br />

The slip angle is the angle between the<br />

wheel’s direction <strong>of</strong> heading (wheel plane)<br />

and its direction <strong>of</strong> travel. So, to compute<br />

the slip angle, need to track tire velocity<br />

components, as shown below.<br />

Excerpt from Liljedahl, et al (1996)<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Additional notes/comments on Ackermann steering<br />

• At low speed the wheels will roll without slip angle.<br />

• If the rear wheels have no slip angle, the center <strong>of</strong> the turn lies<br />

on the projection <strong>of</strong> the rear axle. Each front steered wheel has<br />

a normal to the wheel plane that passes through the same center<br />

<strong>of</strong> the turn. This is what Ackermann geometry dictates.<br />

• Correct Ackermann reduces tire wear and is easy on terrain.<br />

• Ackermann steering geometry leads to steering torques that<br />

increase with steer angle. The driver gets feedback about the<br />

extent to which wheels are turned. With parallel steer, the trend<br />

is different, becoming negative (not desirable in a steering<br />

system – positive feedback).<br />

• Off-tracking <strong>of</strong> the rear wheels, ∆, is related to this geometry.<br />

The ‘∆’ is R[1-cos(L/R)], or approximately L 2 /(2R).<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Using the basic geometry <strong>of</strong> Ackermann steering<br />

L<br />

δ o<br />

B<br />

δ i<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Can you pass the vehicle<br />

through a given position?<br />

1. Assume low-speed turning<br />

2. Project along rear-axle<br />

3. Define R = L/δ max<br />

4. Project from CG<br />

5. Project ideal turning path<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Example: 2D vehicle with front-steered wheel<br />

A wheeled vehicle is said to have kinematic (or Ackermann) steering when a wheel is<br />

actually given a steer angle, δ, as shown. A kinematic model for the steered basic vehicle<br />

in the inertial frame is given by the equations,<br />

‘tricycle’<br />

Xɺ = v cosψ = Rwω<br />

cosψ<br />

Y<br />

δ<br />

Yɺ x<br />

= vsin = R sin<br />

ψ ω ψ<br />

v<br />

ψɺ = tanδ<br />

L<br />

where it is assumed that the wheels do not slip, so we<br />

can control the rotational speed and thus velocity at each<br />

wheel-ground contact.<br />

So, the input ‘control variables’ are velocity, v=R w ω,<br />

and steer angle, δ.<br />

In this example, the CG is located on the rear axle.<br />

These kinematic equations can be readily simulated.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

w<br />

y<br />

Note:<br />

L<br />

1<br />

v = v + v<br />

2<br />

vt<br />

ωz = ψɺ<br />

=<br />

L<br />

vt<br />

tanδ<br />

=<br />

v<br />

( )<br />

1 2<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin<br />

ψ<br />

L = wheel base<br />

X<br />

X


Derivation <strong>of</strong> equations<br />

Y<br />

y<br />

L<br />

ψ<br />

δ<br />

1<br />

v = v + v<br />

2<br />

( )<br />

1 2<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

x<br />

L = wheel base<br />

X<br />

X<br />

Note that the forward velocity at the front wheel<br />

is simply, v, but because <strong>of</strong> kinematic steering<br />

the velocity along the path <strong>of</strong> the wheel must be,<br />

v<br />

ωz = tanδ<br />

L<br />

v δ<br />

=<br />

v<br />

cosδ<br />

This means that the lateral velocity at the front<br />

steered wheel must be,<br />

v = vδ sinδ = v tanδ<br />

Now we can find the angular velocity about the CG, which is located at the<br />

center <strong>of</strong> the rear axle as,<br />

t<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Example: simulation and animation <strong>of</strong> steered tricycle kinematic model<br />

% -----------------------------------------------------<br />

% tricycle_model.m<br />

% revised 2/21/12 rgl<br />

% ----------------------------------------------------function<br />

qdot = tricycle_model(t,q)<br />

global L vc delta_radc delta_max_deg R_w<br />

% L is length between the front wheel axis and rear wheel axis [m]<br />

% vc is speed command<br />

% delta_radc is the steering angle command<br />

% State variables<br />

x = q(1); y = q(2); psi = q(3);<br />

% Control variables<br />

v = vc;<br />

delta = delta_radc;<br />

% kinematic model<br />

xdot = v*cos(psi);<br />

ydot = v*sin(psi);<br />

psidot = v*tan(delta)/L;<br />

qdot = [xdot;ydot;psidot];<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

% Physical parameters <strong>of</strong> the tricycle<br />

L = 2.<strong>04</strong>0; % Length between the front wheel axis and rear wheel axis [m]<br />

B = 1.164; % Distance between the rear wheels [m]<br />

m_max_rpm = 8000; % Motor max speed [rpm]<br />

gratio = 20; % Gear ratio<br />

R_w = 13/39.37; % Radius <strong>of</strong> wheel [m]<br />

% desired turn radius<br />

R_turn = 3*L;<br />

delta_max_rad = L/R_turn; % Maximum steering angle [deg]<br />

R = 6.12 m<br />

δ = 0.33 rad = 19.1 deg<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Summary <strong>of</strong> kinematic vehicle turning<br />

• The models introduced here provide additional review <strong>of</strong><br />

fundamental kinematics principles and how they can be applied<br />

to vehicle systems.<br />

• The concepts <strong>of</strong> differential and Ackermann steering are<br />

demonstrated through simulations.<br />

• The kinematic models are commonly used in mobile robot<br />

applications for path planning, estimation, and control.<br />

• These kinematic models cannot tell you anything about the<br />

effect <strong>of</strong> forces or stability.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Vehicle directional stability<br />

Directional stability refers to a vehicle’s ability to<br />

stabilize its direction <strong>of</strong> motion against disturbances.<br />

So, we are typically concerned with:<br />

•Unforced transient operations (natural response)<br />

•Forced response – aperiodic inputs (step, ramp, etc.)<br />

•Forced response – periodic forcing (sine)<br />

•Steady-state directional response<br />

We’ll discuss both lateral and longitudinal stability.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


2D vehicle <strong>of</strong> Rocard<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

One <strong>of</strong> the earliest analyses <strong>of</strong> vehicle stability was<br />

conducted by Y. Rocard (1954) (Steeds, 1960).<br />

The vehicle was simplified as a rigid rectangular<br />

frame with a wheel at each corner and the plane <strong>of</strong><br />

each wheel is vertical and parallel to the frame.<br />

The steering force is assumed to be directly related to<br />

slip angle,<br />

F = Kα<br />

and the contact forces are assumed not to be affected<br />

by vehicle motion.<br />

There is no steering angle in this model.<br />

First, let’s review the lateral forces using modern<br />

notation.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Lateral tire forces induced by slip angles<br />

The slip angle, α, is derived using body-fixed variables.<br />

α δ<br />

− 1 y 1 z<br />

f = f − tan ⎜ ⎟<br />

vx<br />

α<br />

⎛ l Ω − v ⎞<br />

−1<br />

2 z y<br />

r = tan ⎜ ⎟<br />

vx<br />

⎝ ⎠<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Wong (2001)<br />

⎛ v + l Ω ⎞<br />

⎝ ⎠<br />

From Liljedahl, et al (1996)<br />

v 1 y l<br />

− ⎛ + 1Ωz<br />

⎞<br />

− α f = tan ⎜ ⎟ −δ<br />

f<br />

⎝ vx<br />

⎠<br />

v 1 y l<br />

− ⎛ − 2Ω<br />

z ⎞<br />

− α r = tan ⎜ ⎟ −<br />

�<br />

δ r<br />

⎝ vx<br />

⎠<br />

rear steer<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Tire cornering forces<br />

• The slip angle, α, is the angle between direction <strong>of</strong><br />

heading and direction <strong>of</strong> travel <strong>of</strong> a wheel (OA).<br />

• A lateral force, F yα , also referred to as a cornering<br />

force (camber angle <strong>of</strong> the wheel is zero), is<br />

generated at a tire-surface interface, but this may not<br />

be collinear with the applied force at the wheel<br />

center.<br />

• A torque is induced, referred to as a self-aligning<br />

torque. This torque helps a steered wheel return to<br />

its original position after a turn.<br />

• The distance between these two applied forces is<br />

called the pneumatic trail.<br />

• The self-aligning torque is given by the product <strong>of</strong><br />

the cornering force and the pneumatic trail.<br />

• For more on self-aligning moment and pneumatic<br />

trail, see Wong, Section 1.4.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

t p<br />

Fs<br />

Fyα “…side slip is due to the<br />

lateral elasticity <strong>of</strong> the<br />

tire.”<br />

Fs<br />

Fyα Wong<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Typical data on tire cornering forces<br />

“linear region”<br />

Fig. 1.23 from Wong<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

α<br />

Maximum cornering forces:<br />

•passenger car tires: 18 degrees<br />

•racing car tires: 6 degrees<br />

(Wong)<br />

Fig. 1.24 from Wong<br />

Ratio to normal load<br />

α<br />

Variables that impact cornering force:<br />

•Normal load<br />

•Inflation pressure<br />

•Lateral load transfer<br />

•Size<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Cornering stiffness and coefficient<br />

• The cornering stiffness will depend on tire<br />

properties such as:<br />

– tire size and type (e.g., radial, bias-ply, etc.),<br />

– number <strong>of</strong> plies,<br />

– cord angles,<br />

– wheel width, and<br />

– tread.<br />

• Dependence on load is taken into account<br />

through the cornering coefficient,<br />

where Fz is the vertical load.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

CC<br />

α =<br />

C<br />

α<br />

F<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin<br />

z


Cornering stiffness and coefficients<br />

Cornering stiffness for car, light truck<br />

and heavy truck tires<br />

Fig. 1.27 from Wong<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Effect <strong>of</strong> normal load<br />

on the cornering<br />

coefficient<br />

Fig. 1.28 from Wong<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Now, Rocard’s linearized model<br />

Rocard derived two 2nd order ODEs for this<br />

problem*, and for linear approximations found<br />

the characteristic equation,<br />

( )<br />

2 2<br />

s + Rs + S s = 0<br />

2 2<br />

⎡ ⎤<br />

2 ⎛ a ⎞ ⎛ b ⎞<br />

R = ⎢K1 ⎜1+ K 2 ⎟ + 3 ⎜1+ 2 ⎟⎥<br />

MV ⎣ ⎝ k ⎠ ⎝ k ⎠⎦<br />

2<br />

4 K1K3 ( a + b) 2( K1a − K3b) S = − 2 2<br />

Mk<br />

( MkV )<br />

A critical speed<br />

is defined by:<br />

V<br />

=<br />

2 1 3<br />

c<br />

1 − 3<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

2<br />

2 K K ( a + b)<br />

M ( K a K b)<br />

*Refer to Steeds handout for details; we’ll<br />

come back to this later<br />

Stable for all speeds if:<br />

2<br />

I = Mk<br />

K b > K a<br />

3 1<br />

•choose the position <strong>of</strong> the mass center <strong>of</strong> the<br />

vehicle and the steering force characteristics for<br />

the front and rear tires<br />

•If the steering force are equal, then stability is<br />

assured if b is greater than a, or putting the CG in<br />

front <strong>of</strong> the midpoint <strong>of</strong> the wheelbase.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Directional Stability per Rocard<br />

Let’s match this up with modern notation:<br />

V<br />

2<br />

2<br />

2 2 K1K3 ( a + b)<br />

f α rα<br />

c = =<br />

M K1a − K3b M C f α L1 − Crα L2<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

K = C<br />

2C<br />

C L<br />

( ) ( )<br />

1<br />

K = C<br />

R factor = Cr L2 − C f L1<br />

> 0 (for stability at all speeds)<br />

α α<br />

So, only compute a critical speed if R factor is less than or equal to zero.<br />

A critical speed is defined by,<br />

3<br />

f α<br />

rα<br />

(Steeds, 1960)<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


The ‘Bicycle’ or Single-Track Model<br />

Adapted from Wong<br />

How would you add a disturbance?<br />

Need to determine the ‘external forces’.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

•Assume: symmetric vehicle, no roll or pitch<br />

•Represent the two wheels on the front and rear<br />

axles by a single equivalent wheel.<br />

•The bicycle model will have at least three states:<br />

–forward translational momentum or velocity <strong>of</strong><br />

the CG<br />

–lateral translational momentum or velocity <strong>of</strong> the<br />

CG<br />

–yaw angular momentum or velocity about the CG<br />

Refer to Wong, Chapter 5, Eqs. 5.25 – 5.27:<br />

( ɺ ) �<br />

m vx −VyΩ z = Fxf cos( δ f ) + Fxr − Fyf<br />

sin( δ f )<br />

����� �����<br />

( ɺ )<br />

front drive rear drive lateral force effect<br />

m v + V Ω = F + F cos( δ ) + F sin( δ )<br />

y x z yr yf f xf f<br />

I Ω ɺ = l F cos( δ ) − l F + l F sin( δ )<br />

z z 1 yf f 2 yr 1 xf f<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Z<br />

tan<br />

−1<br />

⎛ y 2 z ⎞<br />

⎜ ⎟ = δ r −α<br />

r<br />

vx<br />

α r<br />

Y<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

x t<br />

v − l Ω<br />

⎝ ⎠<br />

l 2<br />

vy<br />

l 1<br />

y<br />

y t<br />

A more general ‘bicycle’ model schematic<br />

ωz<br />

x<br />

β<br />

vx<br />

α f δf<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin<br />

X<br />

v −1 ⎛ y ⎞<br />

β = sideslip = tan ⎜ ⎟<br />

⎝ vx<br />

⎠<br />

tan<br />

ψ<br />

v + l Ω<br />

− 1 ⎛ y 1 z ⎞<br />

⎜ ⎟ = δ f −α<br />

f<br />

vx<br />

⎝ ⎠


Reduced bicycle model: 2 DOF by assuming V x = constant = V<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

v = V<br />

x<br />

δ = δ = small steering angle<br />

f<br />

( ɺ ) �<br />

m vx −VyΩ z = Fxf cos( δ f ) + Fxr − Fyf<br />

sin( δ f )<br />

����� �����<br />

mvɺ = F + F + F δ − mv Ω<br />

y yr yf xf f x z �<br />

∼0<br />

I zΩ ɺ<br />

z = l1Fyf − l2Fyr + l1Fxf δ f<br />

���<br />

Note how δ falls out, but it enters again in the lateral forces, since,<br />

v 1 y l1 z vy l<br />

− ⎛ + Ω ⎞ ⎛ + 1Ωz<br />

⎞<br />

α f = δ f − tan ⎜ ⎟ ≈ δ f − ⎜ ⎟<br />

⎝ vx ⎠ ⎝ vx<br />

⎠<br />

front drive rear drive lateral force effect<br />

∼0<br />

l 1 2Ω z − vy l2Ω z − v<br />

− y<br />

⎛ ⎞ ⎛ ⎞<br />

αr<br />

= tan ⎜ ⎟ ≈ ⎜ ⎟<br />

⎝ vx ⎠ ⎝ vx<br />

⎠<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


As before, we can calculate vehicle trajectory calculations<br />

Recall, these models solve for forward and lateral velocity and yaw<br />

velocity, resulting from input steer angles, δ, relative to the bodyfixed<br />

axes.<br />

To find the trajectory <strong>of</strong> the CG in the Earth-based coordinates, we<br />

use transformation equations for basic 2-D trajectory simulations,<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Xɺ = v cos( ψ ) − v sin( ψ )<br />

x y<br />

Yɺ = v sin( ψ ) + v cos( ψ )<br />

ψɺ<br />

= Ω<br />

x y<br />

z<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Example: directional stability with zero steer + disturbance<br />

• We examine directional stability, reviewing Steeds’ form <strong>of</strong><br />

Rocard’s (linearized) model (see class handout), which is<br />

essentially the bicycle model with steer angle, δ = 0.<br />

• Determine if this vehicle with baseline parameter data given<br />

(next slide) is stable, and compare with results from a<br />

simulation <strong>of</strong> the bicycle model subjected to a force<br />

‘perturbation’ at the front wheel applied in the +y direction.<br />

• Additional ‘car specification’ data is provided as Appendix B.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Example: baseline vehicle parameter data (from simple_bicycle.m)<br />

(In Matlab script form)<br />

% Baseline values<br />

g = 9.81;<br />

L = 3.075; L1 = 1.568; L2 = L-L1;<br />

% Inertia parameters<br />

m = 1945; % total mass, kg<br />

W = m*g; % weight, N<br />

iyaw = 0.992; % yaw dynamic index<br />

% the following is a defined relation between yaw dynamic index and<br />

% the yaw moment <strong>of</strong> inertia (see Dixon reference and table <strong>of</strong> car specs)<br />

Iz = iyaw*m*L1*L2; % moment <strong>of</strong> inertia about z, kg-m^2<br />

Wf = L2*W/L; % static weight on front axle<br />

Wr = L1*W/L; % static weight on rear axle<br />

% Refer to Wong, Section 1.4 for guide to the following parameters<br />

CCf = 0.171*180/pi; % front corning stiffness coefficient, /rad<br />

CCr = 0.5*0.181*180/pi; % rear corning stiffness coefficient, /rad<br />

Cf = CCf*Wf/2; % corning stiffness per tire, N/rad (front)<br />

Cr = CCr*Wr/2; % rear cornering stiffness per tire, N/rad (rear)<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Disturbance force application<br />

if (t>=tdon & t


For Nominal Case, Vehicle Stabilizes<br />

Given case:<br />

R factor = 4200.4<br />

V c = non-existent<br />

v x =V= 5L/sec<br />

A lateral pulse disturbance is<br />

applied at front wheel.<br />

Afterwards, the vehicle<br />

stabilizes (lateral and yaw<br />

velocities go to zero).<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

4 x 10-3 Lateral Velocity<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

0 5 10 15<br />

50<br />

0<br />

-50<br />

Position<br />

0 50 100 150 200<br />

6 x 10-3 Yaw Velocity<br />

4<br />

2<br />

0<br />

-2<br />

0 5 10 15<br />

x 10-3 Yaw<br />

3<br />

2<br />

1<br />

0<br />

0 5 10 15<br />

Used: simple_bicycle.m<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Change rear ‘lateral stiffness’ (i.e., the cornering stiffness values <strong>of</strong> tires)<br />

In this case, the rear lateral<br />

stiffness was cut in half.<br />

R factor = −33814.<br />

V c = 18.2 m/s<br />

Let (since vehicle is now<br />

unstable for V > Vc)<br />

V = 1.2* V c =21.8<br />

After the pulse disturbance,<br />

the vehicle does not stabilize.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

50<br />

0<br />

-50<br />

Lateral Velocity<br />

-100<br />

0 5 10 15<br />

100<br />

50<br />

0<br />

-50<br />

Position<br />

0 100 200<br />

15<br />

10<br />

5<br />

Yaw Velocity<br />

0<br />

0 5 10 15<br />

15<br />

10<br />

5<br />

Yaw<br />

0<br />

0 5 10 15<br />

Used: simple_bicycle.m<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Example: a double lane change using simple bicycle model<br />

Using the stable case for the<br />

‘simple_bicycle’ model, an<br />

‘open loop’ lane change is<br />

achieved by:<br />

(put right into function file):<br />

% double lane change<br />

if (t=1 & t=2 & t=3 & t=4 & t=5 & t=6) deltaf = 0; end;<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

1<br />

0.5<br />

0<br />

-0.5<br />

Lateral Velocity<br />

-1<br />

0 2 4 6 8<br />

6<br />

4<br />

2<br />

0<br />

Position<br />

-2<br />

0 50 100 150<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

Yaw Velocity<br />

-0.4<br />

0 2 4 6 8<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

Yaw<br />

-0.4<br />

0 2 4 6 8<br />

Used: simple_bicycle_lanechange.m<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Bicycle model summary<br />

• The classic bicycle (or single-track) model forms the basis for<br />

understanding steering and steering control.<br />

• Lateral forces induced by tires depend on wheel slip angle and<br />

play a key role in lateral stability. See also the Rocard handout.<br />

• We saw earlier that longitudinal traction (driving or braking)<br />

influences lateral forces, so driven bicycle models (where there<br />

are traction/braking forces) require us to model the coupling<br />

between longitudinal and lateral tire forces (friction ellipse).<br />

• In extreme maneuvers, for example, it would be possible to<br />

predict yaw instability if lateral forces were significantly<br />

reduced.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

References<br />

1. Den Hartog, J.P., Mechanics, Dover edition.<br />

2. Dixon, J.C., Tires, Suspension and Handling (2nd ed.), SAE, Warrendale, PA,<br />

1996.<br />

3. Greenwood, D.T., Principles <strong>of</strong> Dynamics, Prentice-Hall, 1965.<br />

4. Gillespie, T.D., Fundamentals <strong>of</strong> Vehicle Dynamics, SAE, Warrendale, PA, 1992.<br />

5. Hibbeler, Engineering Mechanics: Dynamics, 9th ed., Prentice-Hall.<br />

6. Meriam, J.L. and L.G. Kraige, Engineering Mechanics: Dynamics (4th ed.), Wiley<br />

and Sons, Inc., NY, 1997.<br />

7. Rocard, Y., “L’instabilite en Mecanique”, Masson et Cie, Paris, 1954.<br />

8. Segel, L., “Theoretical Prediction and Experimental Substantiation <strong>of</strong> the<br />

Response <strong>of</strong> the Automobile to <strong>Steering</strong> Control,” The Institution <strong>of</strong> <strong>Mechanical</strong><br />

Engineers, Proceedings <strong>of</strong> the Automobile Division, No. 7, pp. 310-330, 1956-7.<br />

9. Steeds, W., Mechanics <strong>of</strong> Road <strong>Vehicles</strong>, Iliffe and Sons, Ltd., London, 1960.<br />

10. Wong, J.Y., Theory <strong>of</strong> Ground <strong>Vehicles</strong>, John Wiley and Sons, Inc., New York,<br />

2001 (3rd ed.).<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

APPENDIX<br />

Example Car Specifications*<br />

*Table C.1 from J.C. Dixon, “Tires, Suspension and Handling” (2 nd ed.), SAE, Warrendale, PA, 1996.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Appendix:<br />

More on <strong>Steering</strong> Mechanisms<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


DaNI: 4-wheeled, differentially driven<br />

•Motor: Tetrix DC (handout)<br />

•Gear Ratio: 2:1<br />

•Shaft Diameter: 4.73 mm<br />

•Wheel Diameter: 100 mm<br />

•Wheel base: 133 mm.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Under what conditions would you need to<br />

use a dynamic model?<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


General <strong>Steering</strong> System Requirements<br />

• A steering system should be insensitive to disturbances from the ground/road<br />

while providing the driver/controller with essential ‘feedback’ as needed to<br />

maintain stability.<br />

• The steering system should achieve the required turning geometry. For<br />

example, it may be required to satisfy the Ackermann condition.<br />

• The vehicle should be responsive to steering corrections.<br />

• The orientation <strong>of</strong> the steered wheels with respect to the vehicle should be<br />

maintained in a stable fashion. For example, passenger vehicles require that<br />

the steered wheels automatically return to a straight-ahead stable equilibrium<br />

position.<br />

• It should be possible to achieve reasonable handling without excessive<br />

control input (e.g., a minimum <strong>of</strong> steering wheel turns from one locked<br />

position to the other).<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Passenger <strong>Steering</strong> Requirements<br />

• Driver should alter steering wheel angle to keep deviation from course low.<br />

• Correlation between steering wheel and driving direction is not linear due to:<br />

a) turns <strong>of</strong> the steering wheel, b) steered wheel alterations, c) lateral tire<br />

loads, and d) alteration <strong>of</strong> driving direction.<br />

• Driver must steer to account for compliance in steering system, chassis, etc.,<br />

as well as need to change directions.<br />

• Driver uses visual as well as ‘haptic’ feedback. For example, roll inclination<br />

<strong>of</strong> vehicle body, vibration, and feedback through the steering wheel (effect <strong>of</strong><br />

self-centering torque on wheels).<br />

• It is believed that the feedback from the steering torque coming back up<br />

through the steering system from the wheels is the most important<br />

information used by many drivers.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Ref. Milliken & Milliken<br />

Impact <strong>of</strong> <strong>Steering</strong> Geometry<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

B = 0.56<br />

L<br />

Ref. Wong<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Ref. Wong<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Steering</strong> Results and<br />

Error in Ackermann<br />

B = 0.56<br />

L<br />

“<strong>Steering</strong> error curves” for a front beam axle<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Impact on <strong>Steering</strong> Geometry<br />

<strong>of</strong> using Ackermann in High Speed<br />

A car with a steering geometry chosen for low-speed (Ackermann)<br />

will not be as effective at higher vehicle speeds.<br />

In a high speed turn, the inside tire will have a lower normal force,<br />

which means it could achieve the same amount <strong>of</strong> lateral cornering<br />

force with a smaller slip angle.<br />

Using Ackermann will result in more instances in which the inside<br />

tire is dragged along at too high a slip angle, unnecessarily raising<br />

the temperature and slowing the vehicle down with excessive slip<br />

induced drag.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


1. <strong>Steering</strong> arm<br />

2. Drag link<br />

3. Idler arm<br />

4. Tie rod/rack<br />

5. <strong>Steering</strong> wheel<br />

6. <strong>Steering</strong> shaft<br />

7. <strong>Steering</strong> box<br />

8. Pitman arm<br />

<strong>Steering</strong> Systems – Rigid Axle<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Rack and pinion is not suitable for steering wheels on rigid<br />

front axles, as the axles tend to move in the longitudinal<br />

direction. This movement between the wheels and steering<br />

system can induce unintended steering action.<br />

Only steering gears (1) with rotational movement are used.<br />

Design must minimize effects from motion.<br />

7<br />

8<br />

2<br />

1<br />

You may see some older model<br />

Toyota Land Cruisers that use<br />

this type <strong>of</strong> steering design.<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Heisler (1999)<br />

Axle-Beam <strong>Steering</strong> Linkage<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Heisler (1999)<br />

Axle-Beam <strong>Steering</strong> Linkage<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Front Axle and Tie Rod Assembly<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

From: J.W. Durstine, “The Truck <strong>Steering</strong> System From<br />

Hand Wheel to Road Wheel”, SAE, SP-374, 1973 (L. Ray<br />

Bukendale Lecture).<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


<strong>Steering</strong> Gearbox and Ratio<br />

The gearbox provides the primary means for reducing the rotational input from<br />

the steering wheel and the steering axis.<br />

<strong>Steering</strong> wheel to road wheel ratios may vary<br />

with angle, but have values <strong>of</strong> 15:1 in passenger<br />

cars and may go as high as 36:1 for heavy<br />

trucks.<br />

Rack and pinions are commonly designed to<br />

have a variable gear ratio depending on steer<br />

angle.<br />

The actual steer ratio can be influenced by<br />

steering system effects, such as compliance.<br />

The plot here from Gillespie shows how much<br />

this can change.<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Experimentally measured steering<br />

ratio on a truck.<br />

From Gillespie (Fig. 8.18).<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


3 <strong>Steering</strong> arms<br />

7 Tie rod joints<br />

8 <strong>Steering</strong> rack<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Rack-and-Pinion – 1<br />

Used on most passenger cars and some light trucks,<br />

as well as on some heavier and high speed vehicles.<br />

Used on vehicles with independent suspensions.<br />

<strong>Steering</strong> ratio is ratio <strong>of</strong> pinion revolutions to rack travel.<br />

Some advantages: simple, manufacturing ease,<br />

efficient, minimal backlash, tie rods can be joined<br />

to rack, minimal compliance, compact, eliminate<br />

idler arm and intermediate rod<br />

Some disadvantages: sensitive to impacts, greater<br />

stress in tie rod, since it is efficient you feel<br />

disturbances, size <strong>of</strong> steering angle depends on rack<br />

travel so you have short steering arms and higher<br />

forces throughout, cannot be used on rigid axles<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Heisler (1999)<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Rack-and-Pinion – 2<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


Split track-rod with relay-rod<br />

ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Heisler (1999)<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin


ME 360/390 – Pr<strong>of</strong>. R.G. Longoria<br />

Vehicle System Dynamics and Control<br />

Summary<br />

• <strong>Steering</strong> mechanisms were reviewed briefly<br />

• <strong>Steering</strong> geometry – how do you select steering<br />

system?<br />

• <strong>Steering</strong> control – how does it translate into<br />

steering mechanism design?<br />

• There is a complex relationship to suspension<br />

(see Segel, Gillespie)<br />

<strong>Department</strong> <strong>of</strong> <strong>Mechanical</strong> Engineering<br />

The University <strong>of</strong> Texas at Austin

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!