You are on page 1of 7

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/269293958

Study of wheel slip and traction forces in differential drive robots and slip
avoidance control strategy

Conference Paper · June 2014


DOI: 10.1109/ACC.2014.6859308

CITATIONS READS

8 1,046

3 authors:

Edison Orlando Cobos Torres S. Konduri


Texas A&M University Texas A&M University
9 PUBLICATIONS   31 CITATIONS    14 PUBLICATIONS   167 CITATIONS   

SEE PROFILE SEE PROFILE

Prabhakar Reddy Pagilla


Texas A&M University
198 PUBLICATIONS   2,320 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Design and control of web lateral dynamics View project

Control & Combinatorial Optimization View project

All content following this page was uploaded by Prabhakar Reddy Pagilla on 20 June 2018.

The user has requested enhancement of the downloaded file.


Dynamics and Control of a Differential accelerations are used. When a wheel slips, only a portion of the
applied torque to the wheels is utilized toward the linear motion
Drive Robot With Wheel Slip: of the wheel. Wheel slip can cause considerable deviation from
the actual position of a WMR from its desired position. In particu-
Application to Coordination of lar, this deviation is exacerbated when measurements from
Multiple Robots encoders are used to control the wheel angular position which in
turn is used to control the Cartesian position and orientation of the
WMR. Wheel slip is unavoidable in differential drive robot
Shyamprasad Konduri because a differential velocity between the two driven wheels is
Department of Mechanical Engineering, commanded to achieve WMR orientation in the plane.
There have been several studies related to wheel slip in differ-
Texas A&M University,
ential drive robots [2–9], wherein the kinematics and dynamics of
College Station, TX 77843 the WMR under slip are derived. Wheel slip can occur in the lon-
e-mail: konduri@tamu.edu gitudinal and/or lateral direction of the wheel motion depending
on the wheel acceleration and the traction forces between the
Edison Orlando Cobos Torres wheel and the ground. Traction force expressions are derived for a
Department of Mechanical Engineering, particular type of wheel–ground interaction and they could be
Texas A&M University, complex depending on the wheel and ground properties. Various
combinations of wheel ground interactions have been studied in
College Station, TX 77843
the literature. For instance, a flexible wheel, rigid ground or rigid
e-mail: orlando.cobos@tamu.edu wheel, flexible ground interaction was used in Refs. [2,3], while
other work such as Refs. [4,5], considered rigid wheel, rigid
Prabhakar R. Pagilla1 ground interaction. Some studies employ empirical wheel–ground
Professor interaction force relations that are found based on extensive
Fellow ASME experimentation for various conditions.
Department of Mechanical Engineering, Several studies in the literature have proposed control laws for
differential drive robots with wheel slip. In Ref. [4], a two-stage
Texas A&M University,
controller is employed for trajectory tracking. Traction forces are
College Station, TX 77843 determined by using the static friction coefficient between the
e-mail: ppagilla@tamu.edu wheel and the ground for both slip and nonslip conditions in
Ref. [5]. An optimal model predictive controller was used to con-
trol the differential drive robot when subjected to wheel slippage
In differential drive robots, wheel slip severely affects the ability in Ref. [7]. Regulation and turning controllers were developed in
to track a desired motion trajectory and the problem is exacer- Ref. [6] using a dynamic model that includes lateral and longitudi-
bated when differential drive robots are used in applications nal wheel slip. In Ref. [8], the lateral slip is added as a nonlinear
involving coordination of multiple robots. This problem is investi- perturbation to the system and, in Ref. [9], slip is considered as an
gated and, based on the wheel–ground traction forces, a simple uncertainty and estimated. A majority of the controllers proposed
slip avoidance control strategy is discussed. Differential drive in the literature lack experimental validation. Some models
robots with two driven wheels and one or more ball-type caster involve the use of empirical force models which may require con-
wheels are considered. The traction forces between the wheels siderable testing for individual wheel and surface conditions
and the ground surface are determined by assuming rigid wheel, which may not be relevant once the surface conditions change.
rigid ground interaction. These traction forces are used to deter- Therefore, a simpler and effective method to eliminate the effect
mine the maximum value of the input wheel torque that can be of wheel slip that depends on easily measurable parameters would
applied on the wheel before it slips. To avoid wheel slip, this limit- be useful.
ing torque value is used to set a saturation limit for the input tor- In robot coordination applications where formations are
que computed by a trajectory tracking controller. Stability of the achieved and maintained, the effect of wheel slip is magnified. In
closed-loop system with the slip avoidance strategy is shown. vehicle platoons, a spacing error may propagate along the platoon.
Experiments are conducted with this strategy using a single robot To correct these spacing errors, each robot may have to rapidly
as well as multiple robots in a platoon. A representative sample of accelerate and decelerate leading to wheel slip which in turn may
the experimental results is presented and discussed. exacerbate the spacing errors. Recent work that considered wheel
[DOI: 10.1115/1.4034779] slip in formation control problems may be found in Refs. [10] and
[11]. In Ref. [10], an artificial potential function was used to cre-
ate an adaptive formation control law for systems with lateral slip
1 Introduction
and parameter uncertainties. The effect of wheel slip in a leader
Research related to wheeled mobile robots (WMRs) in a variety follower type formation controller was explored in Ref. [11] to
of areas, such as modeling, control design, and coordination, has show that slip may cause instabilities in otherwise stable
been active for over two decades. Because of their simplicity in formations.
construction and dynamics, differential drive mobile robots have In this work, we study the differential drive robot under wheel
been commonly used. A typical differential drive robot consists of slip along with the forces acting at the wheel ground interaction.
two driven wheels and one or more caster wheels. Most of the The maximum force that can be applied on the wheel before it
research in differential drive robots has assumed pure rolling of slips is determined and used as a saturation limit on the input
wheels, that is, all the torque provided to the wheels is used to wheel torque computed by the controller. Analysis is carried out
impart motion [1]. In practice, the pure rolling assumption does to determine the effect of this saturation limit on the closed-loop
not always hold, especially in cases where large wheel stability of the system using an extension of the classical path fol-
lowing controller from Ref. [1]. Experiments are conducted to
establish that the proposed method is useful in eliminating wheel
1
Corresponding author. slip, thereby improving motion coordination performance. The
Contributed by the Dynamic Systems Division of ASME for publication in the
JOURNAL OF DYNAMIC SYSTEMS, MEASUREMENT, AND CONTROL. Manuscript received
rest of the paper is organized as follows: Mobile robot setup and
December 8, 2015; final manuscript received August 22, 2016; published online some preliminary work are discussed in Sec. 2. A trajectory track-
November 11, 2016. Assoc. Editor: Davide Spinello. ing controller for the differential drive robot along with saturation

Journal of Dynamic Systems, Measurement, and Control JANUARY 2017, Vol. 139 / 014505-1
C 2017 by ASME
Copyright V

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jdsmaa/935719/ on 03/30/2017 Terms of Use: http://www.asme.org/a


effect on avoidance of wheel slip is discussed in Sec. 3. Fmax ¼ ls N (3)
Uncertainty in the coefficient of friction and two methods used to
overcome it are discussed in Sec. 4. Experiments and related dis- where ls is the static friction coefficient and N is the normal
cussion using a single robot and a platoon with three mobile reaction. Once the wheel starts to slip, the amount of force that is
robots are presented in Sec. 5. Section 6 summarizes the work and utilized toward the linear motion of the wheel is given by
draws some conclusions. Flin ¼ lk N, where lk is the kinetic friction coefficient. The differ-
ence between the applied force (Eq. (2)) and the force Flin causes
the wheel to continue slipping. Using these traction forces, if we
limit the applied torques on the wheels such that the reaction force
2 Wheel Slip in Mobile Robots is less than the maximum allowable force Fmax, slip can be
Consider the sketch of a differential drive robot shown in avoided. Thus, the maximum torque that can be applied to the
Fig. 1. Let the position and orientation of the robot center of mass wheel to avoid slip is obtained by equating Eq. (2) to the value
(c) be given by the vector ½xc ; yc ; /T . The center of rotation of the lsN, which results in
two drive wheels are joined together with a rigid link perpendicu-  
l N 2Iwy þ mt r 2
lar to the wheel’s plane of rotation. The length of this link is called smax ¼ s (4)
the wheel base of the robot (2b) and its midpoint is the center of mt r
rotation of robot (o). The distance from centers of rotation and
caster wheel to the center of mass is d and e, respectively. The
mass and the radius of the wheel are m and r, respectively, and the
3 Trajectory Tracking Controller
total mass of the robot is mt. The generalized position vector of
the robot is q ¼ ½xc ; yc ; /; hR ; hL T , where h is the angle subtended The control strategy for the robot consists of two loops one
by the wheel w.r.t. its vertical axis. The control inputs are the nested inside the other. The outer loop controller is used to correct
wheel torques (sR, sL). Let s/ ¼ sin / and c/ ¼ cos /. the position errors and the inner loop controls the individual wheel
The dynamics of the robot in pure rolling (without slip) are velocities. Let the reference trajectory for the robot center of rota-
given by tion be ðxr ; yr ; /r Þ and the corresponding velocities that achieve
this trajectory be denoted by (vxr, xr). If the actual position of the
1 1 robot center of rotation is ðx; y; /Þ, then the absolute position error
v_ x ¼ Fvx ; x_ ¼ sx (1)
m It is defined as

where vx and x are the longitudinal and angular velocities of the e ¼ ½xr  x; yr  y; /r  /T
robot, respectively, Fvx and sx are the longitudinal force and
the angular torque on the robot, respectively. The dynamics of the Let E ¼ ½xe ; ye ; /e T be the position error obtained by transform-
robot when the pure rolling conditions are relaxed and wheels slip ing e into the robot’s reference frame using the coordinate
in both longitudinal and lateral directions are derived in Ref. [12]. transformation
For a wheel that is moving in the longitudinal direction under 2 3 2 32 3
pure rolling, assuming that each wheel of the robot bears half the xe cos / sin / 0 xr  x
total weight of the robot, the traction force on the wheel (Fwheel) E ¼ 4 ye 5 ¼ 4 sin / cos / 0 54 yr  y 5 (5)
for the applied shaft torque s are related by the following /e 0 0 1 /r  /
equation:
Differentiating Eq. (5) gives the error dynamics
smt r
Fwheel ¼ (2) 2 3 2 3
2Iwy þ mt r 2 x_ e xd ye  vd þ vxr cos /e
E_ ¼ 4 y_ e 5 ¼ 4 xd þ vxr sin /e 5 (6)
The maximum force that has to be applied on the free wheel /_ e xr  xd
before it starts to slip is
We will prove the stability of the outer loop and the inner loop
separately starting with the proof of the outer loop. Consider the
following well-known trajectory tracking control law for the outer
loop [1]:
vd ¼ vxr cos /e þ kx xe (7a)
xd ¼ xr þ vxr ðky ye þ k/ sin /e Þ (7b)

where d represents the desired velocity and kx ; ky ; k/ are positive


controller gains. We wish to show that ðxe ; ye ; /e Þ ¼ ð0; 0; 0Þ is a
stable equilibrium of the system (6) when using the control law
(7) for all vxr > 0. Let
1 2 
V¼ xe þ y2e þ ð1  cos /e Þ=ky
2
be a candidate Lyapunov function. V is positive for all nonzero
errors (i.e., V > 0 for all E 6¼ 0) and zero only when error is zero
(V ¼ 0 () E ¼ 0). The derivative of V is

V_ ¼ kx x2e  vr k/ sin2 /e =ky

V_ is negative definite around the equilibrium point, i.e., V_ < 0 for


all ððjxe j; /e Þ 6¼ ð0; npÞ; n ¼ 0; 61; 62; :::Þ. Convergence of ye to
Fig. 1 Two wheeled differential drive robot zero is also guaranteed by LaSalle’s invariance principle. Hence,

014505-2 / Vol. 139, JANUARY 2017 Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jdsmaa/935719/ on 03/30/2017 Terms of Use: http://www.asme.org/a


9
E ¼ 0 is a uniformly asymptotically stable equilibrium point of 1
v_ x ¼ ðsL þ sR Þ >
=
E_ ¼ 0. mr (13)
Let the inner loop controller be as follows: b >
ð 9 x_ ¼ ðsL  sR Þ ;
rI
Fvx ¼ mv_ d þ kp ve þ ki ve dt >>
=
ð (8) We will show that, for longitudinal motion (x ¼ 0), the system is
> stable with saturation on input (sR and sL). With this dynamics,
sx ¼ I x_ d þ kp xe þ ki xe dt >
;
the corresponding matrices used in Eq. (12) are
     
where kp and ki are positive gains. To show the stability of the 0 0 1=mr 1=mr 0
A¼ ; B¼ ; d¼
inner loop, consider the error in the linear and angular velocities: 0 0 b=rI b=rI 0
ve ¼ vd  vx ; xe ¼ xd  x: Differentiating the linear velocity
error and substituting v from Eq. (1) and Fvx from Eq. (8), we In order to prove the stability of the system with saturated input,
obtain one has to show that the pair (A, B) is stabilizable, A has no posi-
 ð  tive eigenvalues, and the control inputs (11) are bounded. Let
1  
v_ e ¼  kp ve þ ki ve dt (9) k k12
m K ¼ 11
k21 k22
Differentiating again, we obtain
Then
v e þ kp v_ e þ ki ve ¼ 0
m€ (10) 2 3
ðk11 þ k21 Þ ðk12 þ k22 Þ
6 7
Therefore, the gains kp and ki may be selected to obtain the A þ BK ¼ 6 mr
4 bðk11 þ k21 Þ
mr 7
bðk12 þ k22 Þ 5
desired velocity error response. Similarly, we can show that 
rI rI
€ e þ kp x_ e þ ki xe ¼ 0
Ix
The characteristic equation of ðA þ BKÞ is
Notice that, in the inner loop controller, the terms mv_ d and I x_ d  
are the feed-forward terms added to the output of a simple 2 b 1
k þk ðk12 þ k22 Þ  ðk11 þ k21 Þ
proportional-integral controller. The control inputs from Eqs. (7) rI mr
and (8) are all implementable as they depend only on the 2b
measured and reference variables.  2 ðk12 þ k22 Þðk11 þ k21 Þ ¼ 0 (14)
mr I
The individual motor torques are calculated from Fvx and sx as
    The roots of the above equation have nonpositive real parts if k11
r sx r sx and k21 are chosen such that
sR ¼ Fvx þ ; sL ¼ Fvx  (11)
2 b 2 b
ðk11 þ k21 Þ  0 (15)
A saturation limit using Eq. (4) is applied on the input torques in
the above equation. The block diagram of the entire closed-loop Therefore, the pair (A, B) is stabilizable. Hence, the conditions for
system is shown in Fig. 2. In the preceding, we have shown that Lemma 1 are satisfied, so there exists a bounded smooth function
the closed-loop system is stable using the control laws from aðxÞ which makes system (1) LES and its states v and x satisfy
Eqs. (7) and (8) without input saturation. It is yet to be shown that an asymptotic bound from Rn with linear gain and nonzero
the system is stable with the addition of saturation. The following restriction. We still need to show that for a smooth and bounded
Lemma from Ref. [13] is useful in proving local exponential sta- reference trajectory the linear and angular control torques Fvx and
bility for the system (1) with saturation. sw are smooth and bounded.
LEMMA 1. Consider the system Since the velocity errors ve and xe are bounded and exponen-
tially decreasing functions, the integrals of those errors are also
x_ ¼ Ax þ BrðuÞ þ d (12) bounded. Differentiating vd from Eq. (7), we obtain

where x is the state vector, A and B are system and input matrices, v_ d ¼ v_ xr cos /e  vxr sin /_ e þ kx x_ e
respectively, d is an external disturbance, and r is a saturation
function. Suppose (A, B) is stabilizable and A has no eigenvalues Since all the components of v_ d are bounded, v_ d is also bounded.
with strictly positive real parts, then there exists a smooth function Hence, Fvx is bounded and the gains kx, kp, and ki can be chosen
a(x) and a positive real number l such that jaðxÞj  ljxj, and the such that the bound is within the saturation value. Along similar
system given by Eq. (12) with u ¼ aðxÞ is zero-input locally expo- lines, we can also show that sx is bounded. Notice that we have
nentially stable (LES), and its state satisfies an asymptotic bound used the linearized no-slip dynamics in proving the stability with
from Rn with linear gain and nonzero restriction. saturation. This is applicable because the dynamics do not include
Considering the dynamics in Eq. (1), for strictly longitudinal slip as long as the input wheel torques stay below the maximum
motion, written in terms of the wheel torques, we have allowable torque (smax) obtained in Eq. (4).

Fig. 2 Two-loop trajectory tracking controller

Journal of Dynamic Systems, Measurement, and Control JANUARY 2017, Vol. 139 / 014505-3

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jdsmaa/935719/ on 03/30/2017 Terms of Use: http://www.asme.org/a


required in addition to the wheel encoders. This is not always pos-
sible and in such situations encoder measurements can be used to
get rough estimates, as long as the applied input torque is well
below the maximum allowable torque (since s close to smax may
cause the wheels to slip). The relation between input torque s,
wheel rotational acceleration a, and friction coefficient ls is
given by
s ¼ Iwy a þ rls N (16)

The maximum allowable torque (smax) corresponds to a maximum


allowable acceleration (amax) which can be computed using the
above relation. Similarly, one-half of smax also corresponds to an
acceleration which can be computed from measurements using
the wheel encoders. Hence, the relation (16) can be used to verify
if the estimated value of l used is acceptable and can be adjusted
in the event it is not. In this paper, we have used the first approach
of utilizing a value for l that is smaller than the measured value to
Fig. 3 Experimental values of coefficient of friction: (a) ls and
obtain torque saturation limit.
(b) lk

5 Experiments Using Single Robot and Platoons


4 Coefficient of Friction l Each wheeled mobile robot has two independent wheels and a
The maximum allowable torque calculated in Eq. (4) depends free-ball caster wheel, Fig. 4. Each independent wheel is fixed to
on the coefficient of static friction. In the current work, coeffi- the shaft of the motor rated at 12 V, with a no-load speed of 350
cients of friction between the wheel and ground pair for the analy- rotations per minute and a stall torque of 0.78 mN. A pair of quad-
sis are obtained experimentally. The robot is pulled on the desired rature encoders with 1856 counts per revolution mounted on the
surface and the forces required to start moving the wheels and to shaft of the motor is used to measure the speed of the wheels and
keep the wheels moving are measured. These measured forces and calculate the position of the robot. The robot control algorithms
calculated normal reaction force (N) on the wheels are used to are implemented using an Arduino microcontroller along with
compute the coefficients. Multiple trails are conducted in different motor drivers and “Xbee Series 1” communication module and a
areas of the surface. The resulting values of static and kinetic desktop computer containing a Xbee Series 1 as receiver.
coefficient of friction obtained are shown in Fig. 3. The mean The robot parameter values are 2b ¼ 0.21 m, e ¼ 0.095 m,
value of static and kinetic friction coefficients thus obtained are d ¼ 0.055 m, g ¼ 9.81 m/s, h ¼ 0.0216 m, mr ¼ 1.5 kg,
ls ¼ 0.241 and lk ¼ 0.239, respectively. Since the two values are Ir ¼ 0.009753 kg/m2, Iwz ¼ 0.000584 kg/m2, Iwy ¼ 0.001168 kg/m2,
very close to each other and the measurement process is prone to r ¼ 0.0365 m, and mw ¼ 0.064 kg. Using these numerical values,
uncertainties for all practical purposes, both values are taken to be the maximum allowable torque on each wheel is found to be
equal to 0.24. 0.095 mN which is equivalent to an acceleration of approximately
The value of l varies considerably for a wheel and surface pair 45 rad/s2; this is applied as a saturation between the computed
depending on the operating conditions. Environmental factors like input torque and the drive motors, see Fig. 2. Model simulations
dust and spills can cause these variations. If these uncertainties and experiments are conducted to verify the dynamics and the
are not accounted for in the experiments, they may produce dis- traction force relations and are shown in Ref. [12].
crepancies in the results. The following are two methods consid- To test the effect of saturation, the wheels are accelerated for
ered to account for these uncertainties while performing the about 0.5 s using a constant input torque after which the torque is
experiments: (1) One can choose a value for l that is smaller than set to zero (i.e., constant velocity). The position of the robot is cal-
the measured or expected value of l to account for the uncer- culated using the encoder readings and the actual position of the
tainty. For instance, if the measurements of l has a 20% uncer- robot is obtained using the overhead video camera. The maximum
tainty, then using a conservative value of 0.8lo (with lo being the available motor torque of 0.13 mN is used to accelerate the robot.
base value) for experiments will guarantee that irrespective of The positions of the robot from both encoders and video with and
what l is within the uncertain region the wheels do not slip. How- without saturation on the torque input are shown in Fig. 5(a).
ever, such an estimate further limits the range of torques that can When input torque is not saturated, the encoders overestimate the
be used. (2) Another strategy is to obtain an online estimate of l robot position; this is because the encoders measure the angular
using velocity and position measurements. Generally, to accu- displacement and do not account for wheel slip. When the input
rately estimate the friction coefficient, position sensors are torque is restricted to less than 0.095 mN, the position obtained

Fig. 4 Picture of mobile robots in platoon

014505-4 / Vol. 139, JANUARY 2017 Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jdsmaa/935719/ on 03/30/2017 Terms of Use: http://www.asme.org/a


Fig. 5 Evolution of the position of the robot: (a) single robot from encoder and video without and with limits on torque input,
platoon of robots without and with limit on torque input, (b) x-position, and (c) y-position

from both encoder and video match each other. Such a torque lim- follows the desired velocity profile. Notice that, when input tor-
iting strategy will be applied in the coordination experiments ques are limited, the positions change more smoothly than without
next. such torque limits, in particular during accelerations and decelera-
Coordination experiments are conducted with three robots in a tions. Also observe that, when limiting input wheel torque, the
platoon with and without the slip avoidance strategy for each intervehicular spacings are maintained better. The evolution of
robot. The reference position of the robot qref is obtained by the y-positions for both cases is shown in Fig. 5(c). Observe that the
combination of the neighbor vehicles’ position qi1 and the error in the y-position is much smaller when the torque limiting
desired intervehicular spacing di as qref ¼ qi1  di . The commu- strategy is employed. The final positions of the three robots from
nication between the robots follows a ring structure. The robot encoder and manual measurement without limiting torque are
positions throughout the experiments are obtained from the (4.78, 0.02), (4.17, 0.04), (3.55, 0.03) and (4.87, 0.24), (4.26,
encoders and the final positions are measured manually. 0.18), (3.65, 0.42). The x-positions are within 0.1 m, whereas the
The experiment started with zero intervehicular spacing error y-position has larger error. The final positions with slip avoidance
and the coordination controller started after 3 s in the experiment. are (4.97, 0.003), (4.36, 0.005), (3.75, 0.002) and (5.02, 0.07),
The starting position for robots are (0.6096, 0), (0,0), and (4.38, 0.06), (3.77, 0.06); clearly, the errors in both x and y are
(0.6096, 0) meters with 0 deg orientation. The desired interve- much smaller. When input torques are saturated using smax, wheel
hicular spacing is set to 0.6096 m. The velocity profile used slip is avoided thereby improving the accuracy of position
for the experiments is shown in Fig. 6. The evolution of the measurement and reducing the spacing errors.
x-position of the robots without and with limiting the torque is
shown in Fig. 5(b). Notice that the rate of change of position
6 Conclusion
In this work, we have investigated wheel slip and its effect on
coordination of multiple mobile robots. The traction forces
between the wheel and the ground are used to create a saturation
limit on the input wheel torques to avoid slip. Simulation and
experimental results involving single and multiple robots in coor-
dination are used to show the effect of such a saturation. The anal-
ysis in Sec. 3 shows that input saturation does not effect the
stability of the closed-loop system with an appropriate choice of
the controller gains. From the experimental results, it can be
observed that a simple saturation on the input wheel torques is
effective in increasing the reliability of measurements from the
encoders. Future work will include the online estimation of fric-
tion coefficients using an additional position sensor. The current
work mostly involved straight line trajectories and future work
will involve using curvilinear paths which give rise to lateral
Fig. 6 Velocity profile used in experiments wheel slip.

Journal of Dynamic Systems, Measurement, and Control JANUARY 2017, Vol. 139 / 014505-5

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jdsmaa/935719/ on 03/30/2017 Terms of Use: http://www.asme.org/a


References [8] Burghi, T. B., Iossaqui, J. G., and Camino, J. F., 2015, “Stability Analysis of
Perturbed Nonlinear Systems Applied to the Tracking Control of a Wheeled
[1] Kanayama, Y., Kimura, Y., Miyazaki, F., and Noguchi, T., 1990, “A Stable Mobile Robot Under Lateral Slip,” XVII International Symposium on Dynamic
Tracking Control Method for an Autonomous Mobile Robot,” IEEE Interna- Problems of Mechanics, Natal, Brazil, Feb. 22–27.
tional Conference on Robotics and Automation, Vol. 1, pp. 384–389. [9] Hoang, N.-B., and Kang, H.-J., 2014, “An Adaptive Tracking Controller for
[2] Balakrishna, R., and Ghosal, A., 1995, “Modeling of Slip for Wheeled Mobile Differential Wheeled Mobile Robots With Unknown Wheel Slips,” Intelligent
Robots,” IEEE Trans. Rob. Autom., 11(1), pp. 126–132. Computing Theory (Lecture Notes in Computer Science), Vol. 8588,
[3] Tian, Y., Sidek, N., and Sarkar, N., 2009, “Modeling and Control of a Nonholo- D.-S. Huang, V. Bevilacqua, and P. Premaratne, eds., Springer International
nomic Wheeled Mobile Robot With Wheel Slip Dynamics,” IEEE Symposium Publishing, Berlin, pp. 277–284.
on Computational Intelligence in Control and Automation, pp. 7–14. [10] Ze-Su, C., Jie, Z., and Jian, C., 2012, “Formation Control and Obstacle Avoid-
[4] Albagul, A., and Wahyudi, 2004, “Dynamic Modeling and Adaptive Traction ance for Multiple Robots Subject to Wheel-Slip,” Int. J. Adv. Rob. Syst., 9(5),
Control for Mobile Robots,” 30th Annual Conference of IEEE Industrial Elec- p. 188.
tronics Society, Vol. 1, pp. 614–620. [11] Tian, Y., and Sarkar, N., 2012, “Formation Control of Mobile Robots Subject to
[5] Nandy, S., Shome, S., Somani, R., Tanmay, T., Chakraborty, G., and Kumar, Wheel Slip,” IEEE International Conference on Robotics and Automation
C., 2011, “Detailed Slip Dynamics for Nonholonomic Mobile Robotic System,” (ICRA), pp. 4553–4558.
International Conference on Mechatronics and Automation, pp. 519–524. [12] Torres, E. C., Konduri, S., and Pagilla, P., 2014, “Study of Wheel Slip and
[6] Tian, Y., and Sarkar, N., 2014, “Control of a Mobile Robot Subject to Wheel Traction Forces in Differential Drive Robots and Slip Avoidance Control Strat-
Slip,” J. Intell. Rob. Syst., 74(3–4), pp. 915–929. egy,” American Control Conference (ACC), pp. 3231–3236.
[7] Wei, S., Uthaichana, K., Zefran, M., and Decarlo, R., 2013, “Hybrid Model Pre- [13] Teel, A. R., 1996, “A Nonlinear Small Gain Theorem for the Analysis
dictive Control for the Stabilization of Wheeled Mobile Robots Subject to of Control Systems With Saturation,” IEEE Trans. Autom. Control, 41(9),
Wheel Slippage,” IEEE Trans. Control Syst. Technol., 21(6), pp. 2181–2193. pp. 1256–1270.

014505-6 / Vol. 139, JANUARY 2017 Transactions of the ASME

DownloadedViewFrom:
publicationhttp://dynamicsystems.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jdsmaa/935719/
stats on 03/30/2017 Terms of Use: http://www.asme.org/a

You might also like