Open access

Virtual Reality to Simulate Adaptive Walking in Unstructured Terrains for Multi-Legged Robots

Written By

Umar Asif

Submitted: 24 November 2011 Published: 05 September 2012

DOI: 10.5772/46404

From the Edited Volume

Virtual Reality - Human Computer Interaction

Edited by Xin‐Xing Tang

Chapter metrics overview

3,339 Chapter Downloads

View Full Metrics

1. Introduction

Walking robots posses capabilities to perform adaptive locomotion and motion planning over unstructured terrains and therefore are in the focus of high interest to the researchers all over the world. However, true adaptive locomotion in difficult terrains involves various problems such as foot-planning, path-planning, gait dynamic stability and autonomous navigation in the presence of external disturbances [1]. There is a large volume of research work that is dedicated to the development of walking robots that exhibit biologically inspired locomotion. In this context, an insect-sized flying robot [2], an ape-like robot [3], an amphibious snake-like robot ACM-R5 [4], for-legged robots like Big Dog and Little Dog, an ant-inspired robot ANTON [5] and LAURON [6], are some well-known examples of bio-mimetic robots. A related study [7] describes an implementation of a neurobiological leg control system on a hexapod robot for improving the navigation of the robot over uneven terrains.

However, the crucial problem that the general approaches suffer is the challenging task of achieving accurate transition between different gaits while maintaining stable locomotion in the presence of external disturbances [1]. In natural rough terrains, prior knowledge of the terrain information is not always available therefore, this chapter aims to make use of virtual reality technology to inspect adaptive locomotion over unknown irregular terrains for possible improvement in gait mobility and stability in the presence of unexpected disturbances.

In order to evaluate and inspect the dynamic characteristics of an adaptive locomotion over an unstructured terrain, it is difficult and time consuming to conduct real-world experiments often. Therefore, virtual prototyping technology serves a prime role that results in better design optimization and allows inspection and evaluation of the mathematical models specifically those which are complicated or impractical to estimate using real prototypes in real-world [16-18]. Virtual Reality Technology provides a rapid method of validating several characteristics of a mathematical model using dynamic simulations for the identification of potential errors in early design stages [8]. Possible defects can be identified and removed by modifying the simulation model itself which is less time consuming and practical if to be performed often. Thus, the complete control algorithm can be analyzed using the simulation model [8].

The chapter is structured as follows: second section describes the mathematical modelling of a hexapod robot using kinematics and dynamics equations. In the third section, gait generation is described for planning cyclic wave and ripple gaits. Fourth section is dedicated to the gait analysis over flat and uneven terrains. Finally, the chapter draws conclusion on the significance of using a simulation-based approach for rapid-prototyping of gait generation methods in multi-legged robots.

Advertisement

2. Mathematical modeling and design

The studies conducted in [8] & [9] use SimMechanics software to construct a simulation-based model of a six-legged robot which comprises of several subsystems interconnected to each other as described in Fig. 1. The physical robot is represented using a subsystem known as “plant”. Another subsystem which represents the kinematic engine contains the forward and reverse kinematics of the robot. Other subsystems include a gait generation block and an environment sensing block as shown in Fig. 1. The aforesaid subsystems are further described in the following subsections.

<xref rid="F1" ref-type="fig">Figure 1</xref>.

Overall control loop of the simulation model.

2.1. Plant subsystem

The “plant” is shown in Fig. 2 in the form of a block diagram that provides complete information about the mass, motion axes and moments of inertia of the bodies and their subassemblies. Each leg attached to the main body is considered as a 3-degree of freedom serial manipulator which acts as the actuating element in the overall simulation model. Furthermore, the block diagram shows joints to describe appropriate connections between the rigid bodies which provide an opportunity for actuation and sensing. Fig. 3 & Fig. 4 show the respective subsystems of a hexapod robot as investigated in [8, 9].

<xref rid="F2" ref-type="fig">Figure 2</xref>.

A six-legged hexapod robot and its leg kinematic representation.

<xref rid="F3" ref-type="fig">Figure 3</xref>.

SimMechanics model representing the plant subsystem of a hexapod robot.

<xref rid="F4" ref-type="fig">Figure 4</xref>.

SimMechanics model of an individual leg as a 3-DOF serial manipulator.

2.2. Kinematics engine

The kinematic model of the robot reported here uses forward and inverse kinematic equations of a 3-DOF serial manipulator to control the foothold in 3D-space. The research studies [8 - 14] describe the mathematical modelling of a 3-DOF leg in 3D space in terms of a forward kinematic model. The details are given in the following subsections.

2.2.1. Forward kinematics

Each leg is modelled as a 3-DOF serial manipulator comprised of three members as: coxa, femur and tibia inter-related using pin joints. The joint connecting the body with the coxa (BodyCoxa-joint) is represented byϖc, the joint connecting coxa with the femur (CoxaFemur-joint) is represented byϖf, and the joint connecting the femur with the tibia (FemurTibia-joint) is represented by ϖt. The position of the foothold [Xt Yt Zt]Twith respect to BodyCoxa-joint motion axis is obtained using Denavit Hartenberg convention and expressed as a general homogenous transformation expression as given by (1). Table1 enlists the D-H parameters of the robot leg shown in Fig.5.

<xref rid="F5" ref-type="fig">Figure 5</xref>.

Kinematic Configuration of an individual leg.

iϏiϖiaidi
00000
190ϖclc0
20ϖflf0
30ϖtlt0

Table 1.

D-T parameters of leg shown in Fig. 5

XtYtZt=cosϖc0sinϖclccosϖcsinϖc0-cosϖclcsinϖc01000001×E1
cosϖf-sinϖf0lfcosϖfsinϖfcosϖf0lfsinϖf00100001×cosϖt-sinϖt0ltcosϖtsinϖtcosϖt0ltsinϖt00100001E2
(1)

Solving (1) yields (2) which can be further simplified to (3).

XtYtZt=lccosϖc + lfcosϖccosϖf + ltcosϖccosϖfcosϖt - ltcosϖcsinϖfsinϖtlcsinϖc +lfcosϖfsinϖc + ltcosϖfsinϖccosϖt - ltsinϖcsinϖfsinϖtlfsinϖf+ ltcosϖfsinϖt+ ltcosϖtsinϖf E3
(2)
XtYtZt=cosϖc (lc + lt cosϖf+t + lf cosϖf) sinϖc(lc+ltcosϖf+t+lfcosϖf) ltsinϖf+t+ lfsinϖfE4
(3)

2.2.2. X-Y-Z (roll-pitch-yaw) body rotations

In order to obtain the position of foothold with reference to robot motion centroid, the body rotations in 3D-space must be brought into consideration. Body rotations (Roll-Pitch-Yaw) representation for orientation is given by (4).

Roll=psi (Ϧ)=Body Rotation about X-axisYaw= gamma (ϑ)=Body Rotation about Z-axisPitch= phi (Ϥ)=Body Rotation about Y-axisE5
Ft=Rx,Ϧ×Ry,Ϥ×Rz,ϑE6
(4)

Using general homogenous transformation matrices, the foot location with respect to body coordinate system can be obtained using (4) as given by (5)

FtxFtyFtz=10000cosϦ-sinϦ00sinϦcosϦ00001×cosϤ0sinϤ00100-sinϤ0cosϤ00001×cosϑ-sinϑ00sinϑcosϑ0000100001×TxTyTzE7
(5)
FtxFtyFtz=E8
=Tzsinϑ + TxcosϑcosϤ - TycosϑsinϤTx(cosϦsinϤ + cosϤsinϑsinϦ) + Ty(cosϤcosϦ - sinϑsinϤsinϦ) - TzcosϑsinϦTx(sinϤsinϦ - cosϤcosϦsinϑ) + Ty(cosϤsinϦ + cosϦsinϑsinϤ) + TzcosϑcosϦE9
(6)

Where,

{Tx=LegOffsetX+XtTy=LegOffsetY+YtTz=LegOffsetZ+ZtE10

Thus, the forward kinematic model of the leg can be expressed by (6), also investigated in [13].

2.2.3. Inverse kinematics

In order to obtain the inverse kinematics solution, we solve the general homogenous transformation expression for a 3-DOF serial manipulator in 3D-space given by (7) as (8) & (9).

R=T1×T2×T3E11
(7)
invT1×R=T2×T3E12
(8)
invT2×invT1×R=T3E13
(9)

Where,

R=r11r12r13Pxr21r22r23Pyr31r32r33Pz0001, T1=cosϖc0sinϖclccosϖcsinϖc0-cosϖclcsinϖc01000001E14
T2=cosϖf-sinϖf0lfcosϖfsinϖfcosϖf0lfsinϖf00100001, T3=cosϖt-sinϖt0ltcosϖtsinϖtcosϖt0ltsinϖt00100001E15

Solving (8) yields (10)

cosϖcsinϖc0-lc0010sinϖc-cosϖc000001×r11r12r13Pxr21r22r23Pyr31r32r33Pz0001=E16
=cosϖf-sinϖf0lfcosϖfsinϖfcosϖf0lfsinϖf00100001×cosϖt-sinϖt0ltcosϖtsinϖtcosϖt0ltsinϖt00100001E17
(10)

Comparing r34 elements of both sides of the (10) yields (11).

Pxsinϖc - Pycosϖc=0E18
(11)

Thus, BodyCoxa joint angle ϖc can be determined from (11) as given by (12).

ϖc=ATAN2(Py,Px)E19
(12)

Solving (9) yields (13)

cosϖfsinϖf0-lf-sinϖfcosϖf0000100001×cosϖcsinϖc0-lc0010sinϖc-cosϖc000001×E20
×r11r12r13Pxr21r22r23Pyr31r32r33Pz0001=cosϖt-sinϖt0lt×cosϖtsinϖtcosϖt0lt×sinϖt00100001E21
(13)

Comparing r14,r24 elements of both sides of (13) we get (14) & (15)

cosϖfPxcosϖc - lc + Pysinϖc- lf + Pzsinϖf= ltcosϖtE22
(14)
Pzcosϖf sinϖfPxcosϖc lc + Pysinϖc= ltsinϖtE23
(15)

Taking positive squares of both sides of (14) & (15) and summing them-up yields (16).

[cosϖf×A- lf + Pz×sinϖf]2+[Pz×cosϖf - sinϖf×A]2=lt2E24
(16)

Where,

A=(Pxcosϖc - lc + Pysinϖc)E25

Further solving (16) yields:

(A2+ lf2 + Pz2-lt2)2lf=cosϖf×A+sinϖfPzE26
G=cosϖf×A +sinϖfPzE27
(17)

Where,

G=(A2+ lf2 + Pz2-lt2)2lfE28

Using the trigonometry identities given by (18) & (19), (17) can be further solved to determine CoxaFemur angle as given by (20)

cosϖx×a +sinϖx×b=cE29
(18)
ϖx=ATAN2b,a+ATAN2(a2+b2-c2,c)E30
(19)
ϖf=ATAN2Pz,A+ATAN2(A2+Pz2-G2,G)E31
(20)

Where,

a=A,b=Pz,c=GE32

Dividing (14) by (15) yields (21) which can be further simplified to obtain FemurTibia angle as given by (22)

ltsinϖtltcosϖt=Pzcosϖf - sinϖfPxcosϖc - lc + PysinϖccosϖfPxcosϖc - lc + Pysinϖc- lf + PzsinϖfE33
(21)
ϖt=ATAN2(Pzcoscosϖf-sinsinϖfA,(A×coscosϖf-lf+sinsinϖfPz))E34
(22)

2.3. Robot dynamics

Lagrange-Euler formulation is selected here to derive the dynamics of the robot leg. The Lagrange-Euler equations yield a dynamic expression that can be expressed as given by (23).

Ϣe-JTF=Mqq̦+Hq,q̥+GqE35
(23)

In (23), M(q) represents the mass matrix, H incorporates the centrifugal and Coriolis terms, and G(q) is the gravity matrix. Ϣecontains active joint torques and F is a matrix representing ground contact forces estimated using force sensors embedded in each foot.

2.3.1. Foothold dynamics

The position of the foothold with respect to BodyCoxa joint motion axis is given by (3). Time derivate of (3) yields velocity of the foothold as given by (24) which can be expressed in form of Jacobean as given by (25)

Xt̥Yt̥Zt̥=-(lc×sinϖc+lfcosϖfsinϖc+ltsinϖccosϖf+t)-(lfcosϖcsinϖf+ltcosϖcsinϖf+t)lc×cosϖc+lfcosϖccosϖf+ltcosϖccosϖf+t-(lfsinϖcsinϖf+ltsinϖcsinϖf+t)0lfcosϖf+ltcosϖf+tE36
-(lt×cosϖc× sinϖf+t)-(lt×sinϖc× sinϖf+t)lt×cosϖf+t×ϖ̥cϖ̥fϖ̥tE37
(24)
Xt̥=JDϖ̥tE38
(25)

Time derivate of (25) yields acceleration of the foothold as given by (26)

Xț=JD̥ϖ̥t+JDϖțE39
(26)

2.3.2. Femurtibia joint dynamics

The position of the FemurTibia joint with respect to BodyCoxa joint motion axis is given by (27).

XfYfZf=lc cosϖc+ lf cosϖc cosϖf)(lcsinϖc+lfcosϖfsinϖc)0E40
(27)

Time derivate of (27) yields velocity matrix as given by (28) which can be expressed in terms of Jacobean as (29)

Xf̥Yf̥Zf̥=-(lc sinϖc+ lf sinϖc cosϖf)-(lf cosϖc sinϖf)0lc cosϖc+ lf cosϖc cosϖf-(lf sinϖc sinϖf)0000×ϖ̥cϖ̥fϖ̥tE41
(28)
Xf̥=JDϖ̥fE42
(29)

Time derivate of (29) yields acceleration matrix as given by (30) which can be further expressed in form of Jacobean as given by (31).

Xf̦Yf̦Zf̦=-lc cosϖcϖ̥c2-lf cosϖccosϖfϖ̥c2-lf cosϖccosϖfϖ̥f2-ϖc̦lfcosϖfsinϖc--lc sinϖcϖ̥c2-lf cosϖfsinϖcϖ̥c2-lf sinϖccosϖfϖ̥f2+ϖc̦lfcosϖfcosϖc-0E43
ϖf̦lfcosϖcsinϖf-ϖc̦lcsinϖc+2ϖ̥cϖ̥flfsinϖcsinϖfϖf̦lfsinϖcsinϖf+ϖc̦lccosϖc-2ϖ̥cϖ̥flfcosϖcsinϖf0E44
(30)
Xf̦=JD̥ϖ̥f+JDϖf̦E45
(31)

2.3.3. Coxafemur joint dynamics

The position of the CoxaFemur joint with respect to BodyCoxa joint motion axis is given by (32).

XcYcZc=lccosϖclcsinϖc0E46
(32)

Time derivate of (32) yields velocity matrix as given by (33) which can be expressed in terms of Jacobean as (34).

Xc̥Yc̥Zc̥=-lc sinϖc00lc cosϖc00000×ϖ̥cϖ̥fϖ̥tE47
(33)
Xc̥=JDϖ̥cE48
(34)

Time derivate of (34) yields acceleration matrix (35) which can be expressed in terms of Jacobean as (36).

Xc̦Yc̦Zc̦=-lc cosϖcϖ̥c2-ϖc̦lcsinϖc-lc sinϖcϖ̥c2+ϖc̦lccosϖc0E49
(35)
Xc̦=JD̥ϖ̥c+JDϖc̦E50
(36)

2.4. Control algorithm

Dynamics of a robot as given by (23) can be further described using Lagrange-Euler formulations as given by (37).

Q=Mqq̦+Hq,q̥+GqE51
(37)

Where,

M=i=1n(JMiTmiJMi+12JRiTIi0JRi)E52

The kinetic energy and potential energy of a link can be expressed as given by (38) & (39).

K=12q̥TM(q)q̥E53
(38)
P=i=1nmigTrciE54
(39)

The mass matrix M of the robot leg can be written by assuming the leg as a serial manipulator consisting of three links as given by (40).

M=i=13(JMiTmiJMi+12JRiTIi0JRi)E55
(40)

The moments of inertia for each link (coxa, femur & tibia) of the leg can be determined from (41), (42) & (43).

Ic0=Tc0IccTcTE56
(41)
If0=Tf0IffTfTE57
(42)
It0=Tt0IttTtTE58
(43)

The Jacobean terms of (40) can be determined from expressions (44), (45) & (46).

JMc=-lc sinϖc00lc cosϖc00000,JRc=000000100E59
(44)
JMf=-(lc sinϖc+ lf sinϖc cosϖf)-(lf cosϖc sinϖf)0lc cosϖc+ lf cosϖc cosϖf-(lf sinϖc sinϖf)0000, JRf=000000100E60
(45)
JMt=-(lcsinϖc+lfcosϖf sinϖc+ltsinϖccosϖf+t)-(lfcosϖc sinϖf+ltcosϖcsinϖf+t)lccosϖc+lfcosϖc cosϖf+ltcosϖccosϖf+t-(lfsinϖcsinϖf+ltsinϖc sinϖf+t)0lfcosϖf+ltcosϖf+tE61
-(ltcosϖc sinϖf+t)-(ltsinϖc sinϖf+t)ltcosϖf+t, JRt=000000100E62
(46)

The velocity coupling vector H and the gravitational matrices are given by (47) & (48)

Hijk=j=1nk=1n(Mijqk-12Mjkqi)E63
(47)
Gi=j=1nmjgTJMjiE64
(48)

Using desired time-based trajectories of each joint, the robot can be commanded to follow a desired path with the help of a computed torque control law as given by (49)

Dq(q̦-kDe̥-kpe+Hq,q̥+Gq=QE65
(49)

The error vector “e” is given by (50)

e=q-qdE66
(50)

Thus, the computed control law can be described by (51).

Dqqd̦+Hq,q̥+Gq+Dq(-kDe̥-kPe)=QE67
(51)
Advertisement

3. Gait generation

Gait generation is the main topic of investigation in the field of legged locomotion. Mobility and stability are two important factors which require specific attention while planning a gait for locomotion in natural environments. Generally, the gaits found in nature feature sequential motion of legs and the body. A large volume of work [5-14] deals with biologically-inspired gaits to replicate this sequential motion pattern found in insects into real robots. The gaits investigated here are cyclic periodic continuous gaits (wave gait & ripple gait) as investigated in [10 & 12]. Gait generation occurs in the gait planner shown in Fig. 1, where commanded foothold locations and orientations are combined with gait sequences to generate a sequential pattern of footholds in 3D-space. The computed leg-sequential pattern is submitted to the inverse kinematic model to determine appropriate joint angles to achieve the desired location. The motion of each leg is determined by the mathematical model. The sequential patterns of leg-lifts for a 12-step ripple gait and a 19-step wave gait are shown in Fig. 16 while, Fig. 6 shows a pictorial demonstration of the gait generation in simulation model and the actual prototype robot over a smooth flat surface.

<xref rid="F6" ref-type="fig">Figure 6</xref>.

A comparative representation of wave gait generation in simulation model and actual prototype robot over a smooth flat surface.

SimMechanics provides a diverse library of software sensors to obtain information from the environment. These sensors mainly include rigid-body sensors and joint sensors. As shown in the figure, the sensors subsystem outlines a network of sensors employed at each leg and their respective joints. Sensors attached to the legs are used to obtain the absolute location and orientation of each foothold in 3D-space while the joint sensors provide information about each joint’s linear and angular rates and accelerations. This sensory network acquires the information from each sensor, filters the data to reject noise. Motion planner, uses the feedback to compute corrections between the commanded data and the recorded data. A Butterworth high-pass 6th order digital filter is used to filter the sensor noise.

Advertisement

4. Dynamic simulations

The objective of carrying out dynamic simulations is to test the validity of the mathematical model and perform gait analysis over uneven terrains. In order to realize the first objective, the mathematical model is simulated first in forward dynamics mode and then in inverse dynamics mode. The robot is commanded to walk between two fixed points 2 meters apart. Straight line walking is achieved with an attitude control using GPS and Compass sensors feedback. Fig. 7 show respective views of robot straight-line walking achieved over flat surfaces using Microsoft Robotics Studio [15].

4.1. Test 1 – Forward dynamics mode

In forward dynamics mode, the objective is to analyze the torques produced at the joints provided each joint is actuated with a motion profile consisting of joint rotation, velocity and acceleration. The robot is steered with a periodic cyclic gait. The footholds planned by the gait generator are passed to the kinematics engine and appropriate joint angles are computed in return. The obtained joint angle profile of each joint is differentiated to compute joint rates and accelerations. Finally, the computed joint motion profiles are fed to the joint actuators to simulate the motion. Joint sensors are used to record the actual torque produced during the entire sequential motion. Actual torques data obtained from the simulation and those determined by the dynamic model are plotted as shown by Fig. 8 (D).

<xref rid="F7" ref-type="fig">Figure 7</xref>.

A view of a hexapod robot being steered between two fixed points (2-meter apart) over a flat surface.

4.2. Test 2 – Inverse dynamics mode

In inverse dynamics mode, the objective is to analyze the joint motion profiles provided the joints are actuated using torques. Joint sensors are used to record joint rotations, velocities and accelerations which are then matched with those determined by the dynamic model to study the errors. The comparative representation of actual motion data and desired data is shown by Fig. 8(A), Fig. 8(B) & Fig. 8(C). The close matching of the actual and desired data confirms the validity of inverse model for walking over flat surfaces.

<xref rid="F8" ref-type="fig">Figure 8</xref>.

Comparative representation of sensors feedback and mathematical model results. (A) Motion profile of BodyCoxa joint. (B) Motion profile of CoxaFemur joint. (C) Motion profile of FemurTibia joint. (D) Joint torques.

4.3. Test 3 – Gait analysis over an uneven terrain

The objective of this test is to conduct a gait analysis in terms of robot speed, joint reaction torques and reaction forces at the foothold during robot-environment interactions.

4.3.1. Reaction forces at the foothold and reaction torques at the joints

In order to study an adaptive gait over an uneven terrain, we must first describe the joint reaction torques due to the reaction forces which arise at the footholds during the leg-terrain interactions.

4.3.1.1. Reaction torque at bodycoxa joint

Considering Fig. 9, the reaction torque about the BodyCoxa joint motion axis can be expressed by (52), further explained by (53) & (54).

ϢBodyCoxa=r1×FN+r1×FR+r2×WcoxaNmE68
(52)
r1=FtxiG-Xci̠+FtyiG-Ycj̠+FtziG-Zck̠mr2=COMxiG-Xci̠+COMyiG-Ycj̠+COMziG-Zck̠mFN=Nornmal reaction force vector at the footholdFR=Friction force vector at the footholdWcoxa=Weight of the coxa linkXcYcZc=lccosϖclcsinϖc0=Position of Bodycoxa joint axisE69
(53)
ϢBodyCoxa=FtxiG-Xci̠+FtyiG-Ycj̠+FtziG-Zck̠×FNxi̠+FNxj̠+FNxk̠+E70
FtxiG-Xci̠+FtyiG-Ycj̠+FtziG-Zck̠×FRxi̠+FRxj̠+FRxk̠+E71
COMxiG-Xci̠+COMyiG-Ycj̠+COMziG-Zck̠×Wcoxaj̠E72
(54)

<xref rid="F9" ref-type="fig">Figure 9</xref>.

The total torque about BodyCoxa joint axis is equal to the sum of moments due to weight of the coxa link and moments due to friction force and normal reaction forces acting at the foothold.

4.3.1.2. Reaction torque at coxafemur joint

Considering Fig. 10, the reaction torque about the CoxaFemur joint motion axis can be expressed by (55), further explained in (56) & (57).

ϢCoxaFemur=r3×FN+r3×FR+r4×WfemurNmE73
(55)
r3=FtxiG-Xfi̠+FtyiG-Yfj̠+FtziG-Zfk̠mr4=COMxiG-Xfi̠+COMyiG-Yfj̠+COMziG-Zfk̠mFN=Nornmal reaction force vector at the footholdFR=Friction force vector at the footholdWfemur=Weight of the femur linkXfYfZf=lc cosϖc+ lf cosϖc cosϖf)(lcsinϖc+lfcosϖfsinϖc)0=Position of CoxaFemurjoint axisE74
(56)
ϢCoxaFemur=FtxiG-Xfi̠+FtyiG-Yfj̠+FtziG-Zfk̠×FNxi̠+FNxj̠+FNxk̠+E75
FtxiG-Xfi̠+FtyiG-Yfj̠+FtziG-Zfk̠×FRxi̠+FRxj̠+FRxk̠+E76
COMxiG-Xfi̠+COMyiG-Yfj̠+COMziG-Zfk̠×Wfemurj̠E77
(57)

<xref rid="F10" ref-type="fig">Figure 10</xref>.

The total torque about CoxaFemur joint axis is equal to the sum of moments due to weight of the femur link and moments due to friction force and normal reaction forces acting at the foothold.

4.3.1.3. Reaction torque at femurtibia joint

Considering Fig. 11, the reaction torque about the FemurTibia joint motion axis can be expressed by (58), further explained in (59) & (60).

ϢFemurTibia=r5×FN+r5×FR+r6×WtibiaNmE78
(58)
r5=FtxiG-Xti̠+FtyiG-Ytj̠+FtziG-Ztk̠mr6=COMxiG-Xti̠+COMyiG-Ytj̠+COMziG-Ztk̠mFN=Nornmal reaction force vector at the footholdFR=Friction force vector at the footholdWtibia=Weight of the tibia linkXtYtZt=cosϖc (lc + lt cosϖf+t + lf cosϖf)sinϖc(lc+ltcosϖf+t+lfcosϖf)ltsinϖf+t+ lfsinϖf=Position of FemurTibia joint axisE79
(59)
ϢFemurTibia=FtxiG-Xti̠+FtyiG-Ytj̠+FtziG-Ztk̠×FNxi̠+FNxj̠+FNxk̠+E80
FtxiG-Xti̠+FtyiG-Ytj̠+FtziG-Ztk̠×FRxi̠+FRxj̠+FRxk̠+E81
COMxiG-Xti̠+COMyiG-Ytj̠+COMziG-Ztk̠×Wtibiaj̠E82
(60)

<xref rid="F11" ref-type="fig">Figure 11</xref>.

The total torque about FemurTibia joint axis is equal to the sum of moments due to weight of the tibia link and moments due to friction force and normal reaction forces acting at the foothold.

4.3.2. Gait analysis over an uneven terrain

The uneven terrain as shown in Fig. 13 is modelled in Microsoft Robotics Developer Studio. A terrain is modelled with random regions of depression and elevation so that the gait can be analyzed under the influence of unexpected disturbances. The robot is steered between defined waypoints and the joints torques are investigated with different gait patterns. Fig. 12(A) shows a comparative representation of torques computed by our mathematical model and those obtained through joint sensors provided by the simulation engine. The results shown in Fig. 12 are those obtained when the robot was steered with a fast-speed ripple gait. As evident from the forces plot shown in Fig. 12(B), friction force is close to the normal reaction point at some points and exceeds the normal reaction force at some other points which signifies that the robot experienced significant degrees of slip/drift which walking through those regions of the terrain. As a consequence, the mathematical model failed to determine adequate torques at the joints to counter the drift each foothold is experiencing while the locomotion. Therefore, the fast-speed ripple gait was found inefficient over an unexpected terrain due to the non-uniform distribution of payload about the two sides of the robot body in its support pattern.

<xref rid="F12" ref-type="fig">Figure 12</xref>.

Results obtained when the robot was steered using a ripple gait. (A) Comparative representation of actual torques sensed at the joint and the torques data determined by the mathematical model. (B) Plot of normal reaction forces and friction forces at the footholds.

<xref rid="F13" ref-type="fig">Figure 13</xref>.

A view of the robot straight-line walking between desired waypoints over an unknown irregular terrain.

In the second run, the robot was steered with a slow-speed wave gait in which each leg is lifted whereas the remaining supports the body. This confirms an even sharing of payload on either side of the robot. The results of this simulation run are shown in Fig. 14 and Fig. 15. As evident from the Fig. 15(B) that the friction force data never exceeds the normal reaction force which confirms that the robot never experienced a slip/drift over the same track when using the wave gait. As a consequence, close matching of computed torques and the sensed torques during this simulation run as shown by Fig. 15(A) signifies the validity of our proposed mathematical model to realize adaptive walking over uneven terrains at slow speeds. Additional simulation results are exhibited in Fig. 14 which portrays the track of the centre of gravity of the robot and its speed profile. Fig. 15(C) & Fig. 15(D) show the profiles of body rotation angles and rates about x, y & z-axes. As evident from the Fig. 15(C), the roll and pith rotations of the body are minimal which confirms the validity of our attitude control for maintaining the body flat over an unknown irregular terrain.

<xref rid="F14" ref-type="fig">Figure 14</xref>.

A) 12-step ripple gait leg-lift sequential pattern. (B) 19-step wave gait leg-lift sequential pattern.

<xref rid="F15" ref-type="fig">Figure 15</xref>.

A) Path outlined by the robot when using a wave gait. (B) Velocity profile of the robot during its locomotion.

<xref rid="F16" ref-type="fig">Figure 16</xref>.

Results obtained when the robot was steered using a wave gait. (A) Comparative representation of actual torques sensed at the joint and the torques data determined by the mathematical model. (B) Plot of normal reaction forces and friction forces at the footholds. (C) Robot attitude tracking. (D) Angular velocity profiles of the robot’s body.

Advertisement

5. Conclusions

This chapter outlines a study on signifying the importance of employing virtual reality technology in realizing legged locomotion over unknown irregular terrains. The objective is to emphasize that how virtual reality technology can serve a prime role in the improvement of legged locomotion by performing gait analysis over unstructured terrains which is often time intense and complex to conduct in real-worlds. The objective has been achieved by analyzing an adaptive gait of a six-legged hexapod robot with two unique support patterns namely: ripple and wave. A complete mathematical model of the robot is presented at the start of the chapter to equip the reader with the basic knowledge of forward and reverse kinematic equations used in robot kinematics. Robot dynamics is further described with a computed-torque control technique to incorporate the forces and moments at the feet. To validate the proposed mathematical model, simulation tests are first carried out in a constraint simulation environment using SimMechanics and results of mathematical model are compared with those obtained from on-board sensors. For gait analysis, the robot is steered over an unknown irregular terrain first using a fast-speed ripple gait and then using a slow-speed wave gait. Reaction forces at the footholds and the reaction torques about the joint motion axes in a leg are investigated which reveal that the wave gait is slow but more secure for adaptive locomotion over an unknown irregular terrain. This even sharing of active payload on either side of the robot body in wave gait results in greater stability especially in the regions of possible slip/drift. However, when using a fast speed ripple gait in which the force-distribution on either side of the robot body is not uniform, the support pattern fails to provide enough friction force against the normal reaction force that may prevent the slip/drift. As a consequence, ripple gaits provide high mobility and are effective over flat terrains however, wave gaits are more suited to walking over unknown irregular terrains.

Advertisement

Acknowledgement

The author would like to thank the editor for his detailed and pertinent comments.

References

  1. 1. Asif. U.IqbalJ.Motion“.Planningusing.anImpact-based.HybridControl.forTrajectory.Generationin.AdaptiveWalking”.InternationalJournal.ofAdvanced.RoboticSystems.Vol 8 4 212226 , 2011.
  2. 2. Wood, R.J., “The First Take-off of a Biologically Inspired At-Scale Robotic Insect”, in IEEE Transactions on Robotics, 24 24 2 341347 , April 2008.
  3. 3. Kuhn.D.RommermannM.SauthoffN.GrimmingerF.KirchnerF.Concept“.evaluationof. a.newbiologically.inspiredrobot.LittleApe”.inI. E. E. E. R. S. J.InternationalConference.onIntelligent.RobotsSystems 2009 (IROS 09), 589594 , 10-15 October 2009.
  4. 4. Shumei Yu and Shugen Ma and Bin Li and Yuechao Wang, “An amphibious snake-like robot: Design and motion experiments on ground and in water”, in International Conference on Information and Automation 2009 (ICIA 09), 500505 , 22-24 June 2009.
  5. 5. M. Konyev and F.Palis and Y. Zavgorodniy and A. Melnikov and A. Rudiskiy and A. Telesh and U. Sschmucker and V. Rusin, “Walking Robot Anton: Design, Simulation, Experiments”, in International Conference on Climbing and Walking Robots and the Support Technologies for Mobile Machines 2008 (CLAWAR 2008), 922929 , 8-10 September 2008.
  6. 6. Roennau. A.KerscherT.DillmannR.Design“.Kinematicsof. a.Biologically-InspiredLeg.fora.Six-LeggedWalking.Machine”in.theProc. of a Biologically-Inspired Leg for a Six-Legged Walking Machine”, in 2010 Proc. of the 2010 3rd IEEE RAS & EMBS Int. Conf. on Biomedical Robotics and Biomechatronics, The University of Tokyo, Tokyo, Japan, September 2629 .
  7. 7. Willian.A. Lewinger and Roger D. Quinn, “A Hexapod Walks Over Irregular Terrain Using a Controller Adapted from an Insect’s Nervous System”, 2010 2010 IEEE/RSJ International Conference on Intelligent Robots and Systems, October 18-22, 2010, Taipei, Taiwan.
  8. 8. Asif. U.IqbalJ.RapidPrototyping.ofa.GaitGeneration.Methodusing.RealTime.Hardwarein.LoopSimulation. “.InternationalJournal.ofModeling.SimulationScientificComputing. . I. J. M. S. S.C)”vol. 2 4 393411 , 2011.
  9. 9. Asif. U.IqbalJ.DesignSimulationof. a.BiologicallyInspired.HexapodRobot.usingSim.Mechanics“.Proceedingsof.theI. A. S. T. E. D.InternationalConference.Robotics.Robo 2010 128135 , Phuket, Thailand, November 24- 26, 2010.
  10. 10. Asif. U.IqbalJ.ComparativeA.Studyof.BiologicallyInspired.WalkingGaits.throughWaypoint.Navigation“.Advancesin.MechanicalEngineering”.vol 2011 (2011), Article ID 737403, 9 pages, DOI: 10.1155/2011/737403.
  11. 11. Asif. U.IqbalJ.An“.Approachto.StableWalking.overUneven.Terrainusing. a.ReflexBased.AdaptiveGait”.Journalof.ControlScience.EngineeringVolume. 2011 (2011), Article ID 783741, 12 pages, DOI:10.1155/2011/783741.
  12. 12. Asif. U.IqbalJ.AjmalM.Khan“.KinematicAnalysis.ofPeriodic.ContinuousGaits.fora.Bio-MimeticWalking.Robot”Proceedings.of 2011 2011 IEEE International Symposium on Safety, Security and Rescue Robotics, 8085 , November 1-5, 2011, Kyoto, Japan.
  13. 13. Asif. U.IqbalJ.Motion“.Planningof. a.WalkingRobot.usingAttitude.Guidance”International.Journalof.RoboticsAutomationpp. 4148 27 1 2012. DOI: 10.2316/Journal.206.2012.1.206-3589.
  14. 14. Asif. U.IqbalJ.On“.theImprovement.ofLegged.Locomotionin.DifficultTerrains.usinga.BalanceStabilization.Method”International.Journalof.AdvancedRobotic.SystemsVol. 9 1 113 , 2012.
  15. 15. Kyle Johns, Trevor Taylor, Professional Microsoft Robotics Developer Studio (Wiley publishing, Indianapolis, Indiana, 2008).
  16. 16. Asif. U.IqbalJ.ModelingControl.Simulationof. a.Six-DO. F.motionplatform.fora.FlightSimulator. “.InternationalJournal.ofModeling.Simulation”vol. 31 4 307321 , 2011. DOI:10.2316/Journal.20.2011.4.205-5587.
  17. 17. Asif. U.IqbalJ.ModelingSimulation.MotionCues.Visualizationof. a.Six-DO. F.MotionPlatform.for-ManipulationsMicro.International“.Journalof.IntelligentMechatronics.Robotics. I. J. I. M.R)”vol. 1 3 117 , 2011.
  18. 18. Asif. U.IqbalJ.Robotic“. A.Systemwith. a.HybridMotion.CueingController.forInertia.TensorApproximation.in-Manipulations”Micro.InternationalJournal.ofAdvanced.RoboticSystems.Vol 8 4 235247 , 2011.

Written By

Umar Asif

Submitted: 24 November 2011 Published: 05 September 2012