76
Robust Control Applied to Consistent Rendezvous and Docking by Carlos Andrade Submitted to the Department of Aeronautics and Astronautics in partial fulfillment of the requirements for the degree of Engineer in Aeronautics and Astronautics at the EPF ECOLE D’INGENIEURS September 2008 c Massachusetts Institute of Technology 2008. All rights reserved. Author .............................................................. Department of Aeronautics and Astronautics September 04, 2008 Certified by .......................................................... Dr. Alvar Saenz-Otero Lead Scientist, SPHERES Program Thesis Supervisor Accepted by ......................................................... Dr. Max Cerf EADS Astrium

Robust Control Applied to Consistent Rendezvous and Dockingssl.mit.edu/spheres/library/candrade_thesis.pdf · arti cial intelligence algorithms to advance the eld in Distributed

Embed Size (px)

Citation preview

Robust Control Applied to Consistent Rendezvous

and Docking

by

Carlos Andrade

Submitted to the Department of Aeronautics and Astronauticsin partial fulfillment of the requirements for the degree of

Engineer in Aeronautics and Astronautics

at the

EPF ECOLE D’INGENIEURS

September 2008

c© Massachusetts Institute of Technology 2008. All rights reserved.

Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Department of Aeronautics and Astronautics

September 04, 2008

Certified by. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Dr. Alvar Saenz-Otero

Lead Scientist, SPHERES ProgramThesis Supervisor

Accepted by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Dr. Max Cerf

EADS Astrium

2

Robust Control Applied to Consistent Rendezvous and

Docking

by

Carlos Andrade

Submitted to the Department of Aeronautics and Astronauticson September 04, 2008, in partial fulfillment of the

requirements for the degree ofEngineer in Aeronautics and Astronautics

Abstract

The Synchronized Position Hold, Engage, Reorient Experimental Satellites Program(SPHERES) at the MIT Space Systems Laboratory provides research scientist witha facility to incrementally and iteratively mature estimation, control, autonomy andartificial intelligence algorithms to advance the field in Distributed Satellite Systems.This facility is located aboard the International Space Station, which allows to provestate-of-the-art GNC algorithms in a relevant 6 DOF environment.

The present thesis deals with the analysis, design and synthesis of low-authority con-trol algorithms oriented towards a consistent procedure of rendezvous and docking.This consistency radicates in the fact to archieve a maturity level which guaran-tees that the docking will be performed with a success rate higher than 90 percent.Discrete-Time Nonlinear Adaptive controllers are developed to cancel the frictioneffects in ground testing at the SSL. H∞ controllers are implemented to take intoaccount the uncertainties of the real system. Those two controllers are applied inground and into the ISS respectivelly as a milestone towards a generalized ConsistentDocking Technology.

Thesis Supervisor: Dr. Alvar Saenz-OteroTitle: Lead Scientist, SPHERES Program

3

4

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

e

Neither for you

5

6

Contents

1 Introduction 11

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1.2 Automated Rendezvous and Docking . . . . . . . . . . . . . . . . . . 14

1.3 The Space Systems Laboratory . . . . . . . . . . . . . . . . . . . . . 15

1.4 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.4.1 Space Technology Maturation . . . . . . . . . . . . . . . . . . 16

1.5 The SPHERES Testbed . . . . . . . . . . . . . . . . . . . . . . . . . 17

1.5.1 Internal Systems . . . . . . . . . . . . . . . . . . . . . . . . . 19

1.5.2 Control Design in the SPHERES Satellite . . . . . . . . . . . 20

1.6 Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

1.7 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2 Adaptive Control 23

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.2 Adaptive Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.3 Sliding Mode Control . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.4 Adaptive Control of Nonlinear Systems . . . . . . . . . . . . . . . . . 29

2.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.5.1 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.5.2 Experimental Tests . . . . . . . . . . . . . . . . . . . . . . . . 35

2.5.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

7

3 Robust H∞ Control 41

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.2 Satellite Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.3 H∞ Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.3.1 The H∞ standard problem . . . . . . . . . . . . . . . . . . . . 43

3.3.2 Characterization Theorem for Output Feedback . . . . . . . . 44

3.3.3 Mixed Sensitivity Design . . . . . . . . . . . . . . . . . . . . . 45

3.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

3.4.1 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . 48

3.4.2 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . 51

3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4 Consistent Rendezvous and Docking 55

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.1.1 Consistent Docking . . . . . . . . . . . . . . . . . . . . . . . . 55

4.2 Orbital Assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.3 Proposed Docking Algorithm . . . . . . . . . . . . . . . . . . . . . . . 58

4.4 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.4.1 Docking Results in the Air Table . . . . . . . . . . . . . . . . 61

4.5 Docking Results on the Flat Floor . . . . . . . . . . . . . . . . . . . . 63

4.5.1 Docking Test with the PID Controller . . . . . . . . . . . . . 63

4.5.2 Docking Test with the Adaptive Controller . . . . . . . . . . . 64

4.5.3 Docking Test with the H∞ Controller . . . . . . . . . . . . . . 66

4.5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

5 Conclusions 69

5.0.5 Thesis Summary . . . . . . . . . . . . . . . . . . . . . . . . . 69

5.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

5.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

8

List of Figures

1-1 Soyuz Spacecraft Docking with the ISS . . . . . . . . . . . . . . . . . 12

1-2 APDS Docking System aboard the Space Shuttle Discovery . . . . . . 13

1-3 Mating system aboard the ATV . . . . . . . . . . . . . . . . . . . . . 13

1-4 Subsystems included in an RVD system . . . . . . . . . . . . . . . . . 15

1-5 Discontinuity in complexity, risk, and cost at each TRL . . . . . . . . 17

1-6 The SPHERES Satellites Aboard the ISS . . . . . . . . . . . . . . . . 18

1-7 Systems description of the SPHERES satellite . . . . . . . . . . . . . 19

1-8 SPHERES satellite control loop . . . . . . . . . . . . . . . . . . . . . 20

2-1 SPHERES satellite test in the Flat Float . . . . . . . . . . . . . . . . 23

2-2 MRAC Block Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2-3 Sliding mode controller with step reference input xd = 0.2 . . . . . . . 33

2-4 Sliding mode controller with reference input xd = 0.25 sin(2 π

60t)

. . 34

2-5 Adaptive controller with step reference input xd = 0.2 . . . . . . . . . 35

2-6 Convergence of the adaptive parameters fC and fv . . . . . . . . . . . 36

2-7 Dynamic performance of the current PID controller, xd = 0.5 [m] . . . 37

2-8 Tuning of the Nonlinear Controller . . . . . . . . . . . . . . . . . . . 38

2-9 Comparison of the dynamical performances of the proposed controller

vs. the actual controller . . . . . . . . . . . . . . . . . . . . . . . . . 39

3-1 Standard H∞ Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3-2 Block Diagram for the one DOF controller . . . . . . . . . . . . . . . 46

3-3 S/KS sensitivity optimization . . . . . . . . . . . . . . . . . . . . . . 47

3-4 Magnitude Plot of W1 . . . . . . . . . . . . . . . . . . . . . . . . . . 48

9

3-5 Weighted Sensitivity Test . . . . . . . . . . . . . . . . . . . . . . . . 49

3-6 Performance Comparison. Step Input of 0.5[m] . . . . . . . . . . . . . 50

3-7 Performance Comparison. Step Input of 0.5[m] . . . . . . . . . . . . . 52

4-1 SWARM modules being tested at the MSCF flat flor . . . . . . . . . 57

4-2 SWARM Docking Port . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4-3 Docking to a Fixed target Satellite Facing Backwards . . . . . . . . . 59

4-4 Docking Test with the Adaptive Controller at the Air Table . . . . . 61

4-5 Docking Test with the PID Controller at the Flat Floor . . . . . . . . 63

4-6 Step Input response of the Flat Floor Controller xd=0.5 [m] at the Flat

Floor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

4-7 Docking test of the Adaptive Controller at the Flat Floor . . . . . . . 66

4-8 Docking Test with the H∞ controller. Tracking of the Trajectory . . . 67

10

Chapter 1

Introduction

1.1 Motivation

The Rendezvous and Docking or Berthing of spacecrafts (RVD/B) is considered as

a key technology in of spaceflight spacecraft systems.This technology proved to be

crucial during the Apollo Program to ensure safe and successful missions [5]. The first

manned RVD took place when mission Gemini 8 docked the Agena 8 target vehicle.

This mission almost finishes in a disaster when the attitude of the Agena vehicle

malfunctioned, putting both spacecrafts into a dangerous spin. The Gemini RVD

program required a close sincronization both during launch and flight and constant

human control.

The Russian Space Agency (Rosaviakosmos) waw the first to develop an automated

RVD capability. The first unmaned Cosmos spacecrafts rendezvous took place back

in 1967, during the apogee of the space race. Some of the RVD technology current

available in the unmanned spacecraft Progress, and the manned spacecraft Soyuz

are derived from Cosmos. Among these devices there are inertial measurement units

(IMU), rate gyrometer, accelerometer, rendezvous radar and optical devices for final

docking.

11

Figure 1-1 shows the docking point and the optical devices used by the Soyuz space-

craft during final approach to the ISS. This mechanical device uses a reception cone

and capture socket in the target (tnot seen in the picture) and a spherically suspended

rod with shock attenuation on the chaser.

Figure 1-1: Soyuz Spacecraft Docking with the ISS

NASA has yet to develop a fully autonomous RVD capability. Most of the available

experience is derived from Gemini and Apollo. These maneuvers used rendevous

maneuvers planned on ground with a onboard system capable to target and control

the relative position. Finally, the crew had to take over the system for a manual

control over the final capture maneuvers using visual aids. The Space Shuttle still

uses the crew to perform the final maneuvers based on visual aids and is based on

the help of IMU, rate gyrometer, optical equipment and radar. Figure 1-2 shows the

Androgynous Peripherical Docking System (APDS) used aboard the space shuttle,

which is derived from the Apollo-Soyuz program [5]. The first contact is made be-

tween the three petals mounted o the contact rings of each side. The contact ring is

separated from the structural connection ring by a spring-damper system arranged

as a Stewart Plattform. After capture, the contact ring will be retracted using the

screws of the damper elements. Utility connections for powed, data and fluid transfer

are performed automatically.

12

Figure 1-2: APDS Docking System aboard the Space Shuttle Discovery

One of the ’first’ in spaceflight during 2008 was the automated RVD of the Au-

tonomous Transfer Vehicle (ATV) with the ISS. Figure 1-3 shows clearly the docking

mechanism choice in this vehicle, which is a russian probe/drogue similar to the one

found in the Soyuz/Progress spacecraft. ATV includes a series of innovations both in

hardware and GNC software. It uses global GPS positioning during the early phases

of the RVD maneuver and laser range finding for the final and camera-type optical

sensors for the final approach.

Figure 1-3: Mating system aboard the ATV

13

Yet, despite the high sofistication of its hardware, sensors and GNC algorithms, it

has a human element behind the automated space docking. As a matter of fact, the

coordination required for the maiden flight of the ATV was extraoirdinary, requiring

tthe use of three mission control centers (Tolouse, Houston and Korolev), the GPS

array for navigation and the surveillance of the ISS crew. A tradeoff must be found

between safety of the crew and the reduction of required resources by automation. In

the case of a interplanetary mission such as a Mars landing/return, this will not be

possible for the delay in communications ranging from 4 minutes in the best of the

cases to 41 minutes in the worst. It is clear that a fully autonomous RVD technology

must be developed in order to proceed with planetary manned missions

1.2 Automated Rendezvous and Docking

The RVD should not be seen as a simple subsystem, rather as complex and critical

phase of the flight that is integrated with the other subsystems [3]. Some docking-

specific systems such as GNC and the docking have to coordinate smoothly with other

systems of the vehicle such as propulsion, power, communications, etc. It is clear that

the main application of RVD systems during the early years of this century will be

the ISS docking scenario. The requirements of the constellation program, both for

the Altair lunar lander and the Mars mission will require a mature RVD technolgy.

Some current applications of this technology are:

• Orbital assembly of larger spacecrafts.

• Re-supply of orbital platforms and stations.

• Crew exchange in orbital stations.

• Re-joining an orbiting vehicle using a lander

• Orbital re-fueling and repair of spacecrafts

14

Figure 1-4: Subsystems included in an RVD system

The SPHERES testbed, operated by the Space Systems Laboratory (SSL) at MIT

allows to test these algorithms in a safe and iterative way in ground and aboard the

ISS. By doing so, technologies can be proven in a relevant microgravity environment

while being reusable for an incremental development of GNC algorithms. The present

thesis consist in the design and implementation of GNC algorithms in this testbed.

1.3 The Space Systems Laboratory

The Space Systems Laboratory (SSL) is affiliated with the Department of Aeronautics

and Astronautics at the Massachusetts Institute of Technology. Founded in 1995, the

SSL engages in cutting edge research projects with the goal of directly contributing

to the present and future exploration and development of space. Specific missions

include developing the technology and systems analysis associated with small space-

craft, precision optical systems, and International Space Station technology research

and development. [1]

15

The laboratory encompasses expertise in structural dynamics, control, thermal, space

power, propulsion, MEMS, software development and systems. A major activity in

this laboratory is the development of small spacecraft thruster systems as well as

looking at issues associated with the distribution of function among satellites. In

addition, technology is being developed for spaceflight validation in support of a new

class of space-based telescopes that exploit the physics of interferometry to achieve

dramatic breakthroughs in angular resolution. The objective of the Laboratory is to

explore innovative concepts for the integration of future space systems and to train a

generation of researchers and engineers converging in this field.

1.4 Background

1.4.1 Space Technology Maturation

One of the key aspects in the SPHERES plattform is that allows to a reliable, fault

tolerant and relative environment testbed for the development and maduration of

space technologies. Over two decades ago NASA developed the Technology Readi-

ness Levels (TRL) to determine whether a technology is ready for spaceflight. They

are a systematic measurement system that supports assessments of the maturity of

a particular technology and the consistent comparison of maturity between different

types of technology. A detailed description of TRL can be found in [13].

Figure 1-5 shows a a pictorial representation of how complexity, risk, and cost may

increase for a program if it were to follow each TRL one at a time [18]. It can be seen

that TRLs are not simply to follow because they are nonlinear [18]. It could happen

that some of the levels are skipped because the technology is not available. TRL levels

1 to 4 represent small risk, complexity and cost. Scaling up to TRL 5 represents an

increase in costs. Archiving TRL 7 represents a considerable increase the cost, risk

and complexity. Once TRL 8 is reached, the only quantity that increases is the cost

of the development of the space system.

16

Figure 1-5: Discontinuity in complexity, risk, and cost at each TRL

An important factor considered in the TRL review is that the technology has to be

demonstrated in a relevant environment. In the case of space experiments the choice

is limited to free flyer spacecrafts, space shuttle experiments and the ISS. The use

of a free flyer probe allows to interact with a real environment but is highly risky,

expensive and not accessible in case of a failure. The current Space Shuttle program

focuses in the completion of the ISS rather than scientific exploration.

1.5 The SPHERES Testbed

The MIT Space Systems Laboratory (SSL) developed the Synchronized Position Hold

Engage and Reorient Experimental Satellites (SPHERES) program to incrementally

and iteratively mature algorithms for distribute satellite systems in a microgravity

17

environment. SPHERES is a scientific testbed specifically designed to help develop

algorithms relevant to guidance, navigation and control of spacecraft. It provides a

facility with six degrees of freedom (6 DOF) to evaluate the dynamics of a multiple

satellite system. This facility consist of a constellation of three satellites whose can

interactively communicate, maintain position, flight in formation, keep their position

and move to commanded positions as well as detect faults and run internal diagnos-

tics.

Figure 1-6: The SPHERES Satellites Aboard the ISS

The testbed is specifically designed to operate inside the International Space Station

(ISS) but they have been also successfully tested in the Reduced Gravity Aircraft

(KC-135) and in a 2D environment (3 DOF) on the flat floor at the Marshall Space

Flight Center (MSCF) and on the flat table at the SSL. A total of six satellites have

been built, tested and certified for spaceflight. Three of them are aboard the ISS

and the rest are available for ground tests and technology development prior to the

6 DOF tests aboard the ISS.

During the first year of operations with SPHERES aboard the ISS several types

of algorithms have been successfully tested aboard the International Space Station.

These algorithms include Kalman Filters, PID and glideslope control, docking al-

18

gorithms with cooperative and uncooperative targets and satellite formations. The

current lines of research include adaptive control, estimation, fault detection, intelli-

gent control, computer vision and online path planning, among several others.

1.5.1 Internal Systems

The SPHERES testbed consists of multiple micro satellites which can control their

relative positions and orientations in a 6 DOF environment. A laptop computer on-

board the ISS provided by NASA is used as a ground station to transmit commands

to the SPHERES and record telemetry. The ISS crew members use the laptop to

start tests or change test configurations.

Each satellite is self-contained, with all the necessary subsystems to operate au-

tonomously Power is provided by 16 AA batteries. The propulsion system uses pres-

surized 12 CO2 miniature nozzles and electronically actuated micro-solenoid valves

(pulse-width modulation control). The position and attitude are determined using

inertial measurement units (IMU), which consist of three single axis accelerometers

and three single axis gyroscopes.

Figure 1-7: Systems description of the SPHERES satellite

19

Five ultrasonic beacons situated in the test volume provide global positioning to the

satellites. Two RF channels provide independent inter-satellite communications and

satellite-to-laptop communications. The main microprocessor is a Texas Instrument

C6701 DSP processor. ([15])

1.5.2 Control Design in the SPHERES Satellite

The SPHERES satellite control block diagram is shown in figure 1-8. There are four

blocks to be considered:

Figure 1-8: SPHERES satellite control loop

• Controller: The position control determines the thrust commands required to

reach the target position based on the current state of the satellite. There are

two control algorithms: one for position running at 0.5Hz and other for attitude

running at 1Hz.

• Mixer: The Thruster Mixer is a matrix algorithm that converts the thrust

commands sent by the controller into thruster firing times. As mentioned be-

fore, the solenoid valves are controlled by pulse width modulation. The mixer

determines the firing time of each of the thusters.

• Plant: The satellite is considered as a single point mass. It can be easily

considered as a double integrator plant in the one satellite case.

20

• Estimator: An Extended Kalman Filter (EKF) has been implemented to esti-

mate the position. The satellite’s sensors send signal to the Position and Atti-

tude Determination System (PADS) in order to estimate the position and the

satellite estimates its current position respect to a predefined reference frame.

For more precise technical information about the SPHERES testbed and its

operations, the reader might find useful to read [18], [17], [6], [14] and [16].

1.6 Research Objectives

The primary objectives of this research are summarized below:

• Demonstrate the utility of nonlinear control algorithms to compensante the

friction effects found during ground testing.

• Develop robust linear control algorithms for position control and station keep-

ing, docking and automated rendezvous.

• Integrate these control algorithms to docking maneuvers in ground to prove

assembly of in-orbit multibody spacecraft.

1.7 Thesis Outline

Chapter 2 introduces Nonlinear Adaptive Controllers to improve the accuracy of the

ground testing by compensating the contact friction. This controller is expected to

have a good dynamical performance while avoiding any overshoot and reach a static

error of less than one centimeter in order to perform a succesful docking.

Chapter 3 introduces the basic concepts in the synthesis of robust linear controllers

for position control, consistent docking and formation flight in the SISO case. H∞suboptimal controllers have been developed for position control with relative naviga-

tion by using full state knowledge.

21

Chapter 4 introduces the current concepts of RDV and shows the application of the

previous controllers as a part of this technology in different applications. The Non-

linear Adaptive Controller js applied to the SWARM project for assembly of a multi

body spacecraft. The H∞ controller will serve to show robustness in RVD algorithms

under failure.

22

Chapter 2

Adaptive Control

2.1 Introduction

The objective of this chapter is to develop a reference-model nonlinear adaptive con-

troller to compensate the friction between the air carriage containing the satellite and

the glass surface in the flat table. The air carriage, seen in Figure 2-1, is a device

that creates a laminar flow that provides lubrication between the surfaces. There is a

considerable amount of static friction to overcome during the tests because of surface

imperfections in the glass as well as gravity effects due to the inclination of the table.

Figure 2-1: SPHERES satellite test in the Flat Float

23

2.2 Adaptive Control

In the adaptive control methodology (shown in Figure 2-2) the controller estimates the

uncertain constant or slow-varying plant parameters such as the friction parameters

based on the measured system states and using these estimated parameters in the

control input computation [10, pp. 5-8]

Figure 2-2: MRAC Block Diagram

• The plant in this case represents the dynamics of the system (which is a doule

integrator) and the friction dynamics modeled by a known friction model but

including some unknown parameters.

• The reference model specifies the ideal response of the adaptive control system

to the external command. In the present work the reference model is derived

from the concept of the sliding surface.

• The controller is parametrized by the adjustable friction parameters. It has

tracking convergence, which means that it will have an identical output respect

to the reference model output once the parameters are exactly known.

• The adjustment mechanism or adaptation law adjusts the parameters in the

control law such that the response of the plant under adaptive control becomes

the same of the reference model,thus leading to zero tracking error. In the

proposed controller two friction parameters are estimated.

24

Two approaches were taken for the development of the proposed controller. The first

approach is the Sliding Mode control, which is an easy approach in nonlinear robust

control. However this controller often recurs to high control activity and chattering.

The other approach is the adaptive nonlinear controller, which is derived from the

sliding mode control and allows for a more precise control without chattering. In the

following sections both controllers are introduced and simulated. For practical reasons

only the adaptive nonlinear controller is implemented in the SPHERES satellites.

2.3 Sliding Mode Control

The sliding mode control is a simple approach to robust nonlinear control. It is based

in the transformation of a higher order systems into first order systems. It has been

shown [21, pp. 278-283] that tracking performance with a guaranteed precision can

be archived for these transformed problems in the presence of bounded parameter

inaccuracies while incurring in low control activity.

The particular geometric properties of the satellite allow us to make the assump-

tion that the system is a single-mass point. By doing so, the system representation

can be represented by the equations of movement:

m x = Finternal + Fexternal (2.1)

This can be rewritten for the friction problem as:

x = b u+f(x)

m(2.2)

The components of the state vector x = [x x x] represent the position, velocity

and acceleration along the direction ~x. The mass m = 13.5 [Kg] includes both the

satellite and its air carriage. The control input u represents the thrust applied in the

direction ~x. The control gain b = 1/m is known. The function f(x) represents the

estimate of the friction in the system.

25

In the present work, the Coulomb-Viscous friction model has been chosen. This

model represents well the phenomena at relativelly high velocities and stiction but

does not include presliding displacement. It can be described by the equation

f(x) = fC tanh(x

vo

)+ fvx (2.3)

The first term represents the static friction and the second term the viscous friction.

The term fC is the Coloumb friction level, fv is the coefficient of dynamical friction

and vo is the critical velocity. In the scope of this work vo=0.01 [m/s]. This value

is found recurrently on the literature for the same kind of problems (as seen in [22]

and [12]. There are other models that consider more complexe dynamics such as

the Elasto-Plastic model [4], the LuGre model or the Maxwell Generalized model [8].

However those models are not linearly parametrizable. The parameters fC and fv are

linearly parametrizable.

The friction estimate f is assumed to be bounded by a known function F = f(x, x)

such that:

|f − f | =≤ F (2.4)

where

f = −a1

(fC tanh

(x

vo

)+ fv x

)(2.5)

F = a2

(fC tanh

(x

vo

)+ fv x

)(2.6)

With a1=-0.9 and a2=1.1 are the know boundaries of the estimation error on f .

Define the target state vector as xd and the tracking error vector as x = x− xd. Let

us define a time-varying surface S(t) in the state-space R(n) by the scalar equation

s(x; t) = 0 where:

s (x; t) =

(d

dt+ λ

)x (2.7)

26

where λ is a strictly positive constant The problem of tracking x = xd is equivalent

of that of remaining on the surface S(t) for all t > 0 because s0 represents a linear

differential equation with a unique solution x0

It is shown by [21, pp. 279] that, assuming xd(0) = x(0) the tracking problem is

replaced by a first order stabilization problem in s by chosing the control law such

that ouside S(t):1

2

d

dts2 ≤ −η |s| (2.8)

Where η is a strictly positive constant that reflects the time to reach the boundary

layer starting from the outside. S(t) satisfying 2.8 is called a sliding surface. The

system’s behavoir once in S(t) is called sliding mode. Satisfying 2.8 guarantees

that if xd(0) 6= x(0), the surface S(t) will be reached in a finite time smaller than

|s(t = 0)|/η. [21, pp. 281]

The parameter s can be seen as the distance of the state of the system to the sliding

surface, which means it is an analog to the state error. It can be rewritten from its

original definition in 2.7 as:

s = x− xd + λ ˙x (2.9)

Now, including the dynamics of the system to be controlled the parameter becomes:

s = f + u− xd + λ ˙xd (2.10)

The control law estimate, u, is the best approximation to archieve s = 0 and it is

described by:

u = −f + xd − λ ˙x (2.11)

By choosing the control gain k = F + η large enough, the condition 2.8 can be met

([21],pp. 286). In order to satisfy 2.8 despite of the uncertainty in the plant dynamics

f a discontinuous term is added across the surface such that:

u = u− k sgn(s) (2.12)

27

One of the recurrent problems in sliding mode controller is chattering. One of the

easiest methods to compensate this problem is to smooth the control discontinuity in

a boundary layer of tickness φ [21, pp. 290]. This boundary layer can be implemented

as a constant layer φ=0.1. By doing so, the control law becomes:

u = u− k sat(s/φ) (2.13)

Where sat stands for the saturation of s with respect of the boundary φ. The system

performance is very sensitive to the control bandwith λ. In the present work λ is

considered to represent the neglected time delays in communications. A time delay

of TA = 400ms. is considered to take into account the communications delay between

the computer and the satellite. Taking into account the time delay is a solution

proposed in [21, pp. 302] and can be expressed as:

λ ≤ λA ≈1

3TA≈ 0.833 (2.14)

The algorithm 1 shows the proposed sliding mode controller. It receives as inputs the

state x, the target state xd and the error vector e. The output is the control command

u sent to the satellite. Finally, the values of φ and λ are constants explained earlier

in this section. The values of fc and fv are both constant values set to 0.1, which is a

value found currently in literature [8] [4]. The values of η and must be found either

by numerical optimization or by direct experimentation.

The controller developed in this section simplified because several other factors were

not taken into account. For example, the mass m=13[Kg] is supposed to be fixed,

while in reality the difference between the wet mass and the dry mass is in average

400 g. This helps to keep the mass as constant, otherwise other control gains and

time-varying boundaries must be chosen.

28

Alg. 1 ctrl pos sliding(x,xd, e). A position control algorithm to track a determined target

Input: x = [x x], xd = [xd xd], e = [e e]Output: u

1 begin2 define λ, η =, φ, fc, fv3 f = −0.9 fc tanh(x)− 0.9 fv x;4 F = 1.1 fc tanh(x) + 1.1 fv x;5 s = −e− λ e;6 k = F + η;

7 u = −f − λ e;8 u = (u− k sat(s/φ))/m;9 end

2.4 Adaptive Control of Nonlinear Systems

The adaptive control of nonlinear systems has been exploited increasingly in the last

two decades thanks to the current computational tools and recent findings in nonlinear

dynamics. This class of problems usually satisfy the following assumptions [21, pp

350]:

1. The nonlinear plant dynamics can be linearly parametrized

2. The full state is measurable

3. Nonlinearities can be canceled stably (i.e. without unstable hidden modes or

dynamics) by the control input if the parameters are known.

Once again, the single mass-point hypothesis is considered. The system can be de-

scribed by the second-order systen:

x =fCm

tanh(x

vo

)+fvmx = b u (2.15)

where the parameters fc and fv are unknown constants. The parameter b = 1/m

is a known constant. The objective of the adaptive control is to make the output

29

asymptotically track a desired output xd(t) despite the parameter uncertainty.This

can be rewritten as:

h x = fC tanh(x

vo

)+ fv x = u (2.16)

with h = 1/b = m. Let us define a combined error similar to the sliding surface

defined in 2.7:

s = e+ λ e (2.17)

where e is the position error. The scalar s can be rewritten as:

s = x− xr (2.18)

where xr, the reference value of x, is defined as:

xr = xd − λ e (2.19)

xr = xd − λ e (2.20)

The following control law is proposed [21, pp 352]:

u = h xr − k s+ fC tanh(x

vo

)+ fv x (2.21)

where k is a constant of the same sign as h. If the parameters are all known, this

choice gives exponential converge of s which in turn guarantees the convergence of e

[21, pp. 352] .For the adaptive control, the control law 2.21 is modified to include the

estimated values fC and fv

u = h xr − k s+ fC tanh(x

vo

)+ fv x (2.22)

From [21, pp. 328-329] it is posible to show that the adaptation gains can be repre-

sented by:

˙fc = − γ sgn (h) s tanh

(x

vo

)(2.23)

30

˙fv = − γ sgn (h) s x (2.24)

These adaptation gains verify global tracking convergence of the adaptive control

system. In the algorithm 2 the values of γ and are to be found by experimentation

or numerical optimization. T stands for the control period, which by convention is 2

[s] for position control in the SPHERES satellites.

Alg. 2 ctrl pos adaptive(x, xd, e). A position control algorithm to track a determinedtarget by adaptation of the friction parameters.Input: x, xd, eOutput: u

1 begin2 define λ, γ, k, T ;

3 initialize fC , fv, k, T ;4 yr = xd + λ e;5 yr = λ e;6 s = x− yrdot;

7˙fC = −γ s tanh(x);

8˙fv = −γ s x;

9 fC+ =˙fC T

10 fv+ =˙fv T

11 u = m yr − k s+ fC tanh(x) + fv x12 end

2.5 Results

2.5.1 Simulation

The controller has been tested using the discrete simulator available at the SSL, which

emulates the continuous time dynamics of the satellite while taking into account the

discrete operation of the real plant and takes into account the saturation in its actu-

ators.

31

The dynamics engine of ths simulator has been modified in order to add the non-

linear function tanh(x) described in equation 2.3 in its continuous form. This allows

to introduce the nonlinear dynamics in the simulation and emulate a more precise

environment.This function has been discretized in the controller by using a power

series expansion:

tanh(x) =5∑

n=1

B2n 4n(4n − 1)

(2n)!x2n−1 for|x| < π

2(2.25)

where Bn represents the Bernouilli numbers, whose are handily available in mathe-

matical tables.

Figure 2-3 shows the simulation results of the sliding mode controller for a reference

position xd = 0.2. Figure 2-3 (a) shows the evolution of the position with respect

to time. It can be seen that the desired position is reached after 28 seconds. This

delay can be reduced by increasing the control gain η at the price of increasing the

overshoot. An optimal value of η can be found either by numerical optimization or

direct experimentation. In the framework of the present work the values are found

trhough experimentation.

Therefore the true true values of the friction coefficients fC and fv are unknown

because they vary across the working surface. Even to find approximate values would

require a complex nonlinear optimization setup. In the present work both values were

supposed to be 0.01.

Figure 2-3 (b) shows the evolution of the thrust in the direction x with respect to

time. The evolution of s, the distance from the sliding surface, is included to show

that when the sliding surface stabilizes at s=0, the control activity stops. Figure 2-3

(c) shows the evolution of position with respect of the surface s. It can be seen that

32

! "! #! $!!

!%&

!%"

!%'

!%#

()*+,-.+/0

12.)3)24,-*5.0

(67/8)49,:+.;24.+

,

,

<<=

! "! #! $!!!%"

!!%&

!

!%&

!%"

()*+,-.+/0

>26/+,-?0

(@6A.3,162B)C+

,

,

><

.

! "! #! $!!!%"

!!%&

!

!%&

!%"

!%'

()*+,-.+/0

.!.A6B7/+!12.)3)24

DE2CA3)24F,G!GA6B7/+,74=,12.)3)24

,

,

<<=

.

! "! #! $!!!%"

!!%&

!

!%&

!%"

()*+,-.+/0

.!.A6B7/+!12.)3)24

DE2CA3)24F,G!GA6B7/+,74=,12.)3)24

,

,

<<=

Figure 2-3: Sliding mode controller with step reference input xd = 0.2

when the s surface reaches zero, the state reaches its target position. Finally 2-3 (d)

shows the evolution of the tracking error, which reaches zero when s reaches 0

Figure 2-4 shows the tracking response of this controller for a reference trajectory

xd = 0.25 sin(2 π

60t). It can be seen that the sliding mode controller is not good

enough to guarantee perfect tracking, instead, tracking is guaranteed whitin a pre-

cision ε = φ/λ. The tracking error is small (of the order of 5 cm) considering that

the values for η were be found trhough direct experimentation. Figure 2-5 shows the

response of the adaptive controller with a constant reference input xd = 0.2. It can be

seen that the value is easily reached after 18 seconds. The maximal trhust in an axial

direction is about 0.104 [N] which explains this rise time which could be considered

slow in other applications. This rise time is coherent with previous experimentation.

Indeed, in Figure 2-5 (b) it can be seen that a saturation has been implemented in

33

0 20 40 60−0.4

−0.2

0

0.2

0.4

Time [sec]

Posi

tion

[m/s

]

Tracking Response

xxd

0 20 40 60−0.2

−0.1

0

0.1

0.2

Time [sec]

Forc

e [N

]

Thrust Profile

Fx

s

0 20 40 60−0.4

−0.2

0

0.2

0.4

Time [sec]

s−su

rface−P

ositi

on

Evolution: S−Surface and Position

xxd

s

0 20 40 60−0.04

−0.02

0

0.02

0.04

0.06

Time [sec]

s−su

rface−P

ositi

on

Evolution: S−Surface and Position

xxd

Figure 2-4: Sliding mode controller with reference input xd = 0.25 sin(2 π

60t)

software to simulate the maximum thrust available.

Figure 2-5 (c) shows the evolution of the position with respect to the s value, and it

is clearly seen that it reaches its target value and in Figure 2-5 (d) its tracking error

is zero when the s value reaches zero. Figure 2-6 shows the results of the estimates of

this adaptive controller. It can be seen that both values converge but not to the ex-

pected values (fc = fv = 0.01) fixed in the simulation engine. This happens because

the input signal, being a step input, does not excite all the dynamics of the system

because the signal is not rich enough.

34

0 20 40 600

0.05

0.1

0.15

0.2

Time [sec]

Posi

tion

[m/s

]

Tracking Response

xxd

0 20 40 60−0.2

−0.1

0

0.1

0.2

Time [sec]

Forc

e [N

]

Thrust Profile

Fx

s

0 20 40 60−0.2

−0.1

0

0.1

0.2

Time [sec]

s−su

rface−P

ositi

on

Evolution: S−Surface and Position

xxd

s

0 20 40 60−0.2

−0.1

0

0.1

0.2

Time [sec]

s−su

rface−P

ositi

on

Evolution: S−Surface and Position

es

Figure 2-5: Adaptive controller with step reference input xd = 0.2

2.5.2 Experimental Tests

The standard test in this experimentation is a step impulse input of 0.5 m. To have

a constant set of test a PID controller was used to reach an initial position (xd=0.3

[m] and yd=0.3[m]). Before any comparative study can be done it is necessary to see

the performances of the PID controller to see the result of the friction in this con-

troller. The performances of the PID controller for two different test under the same

conditions are shown in figure 2-7. The gains of these controller have been previously

found by tuning and experimentation.

It has to be menctioned that the gains of the PID controller were optimized for

tracking problems. Therefore the present test is not exaustive in the behaviour of the

tracking properties rather than its impulsive response.

It can be seen in figure 2-7 that the rising time is reasonable considering the low thrust

35

0 20 40 600

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

0.045

Time [sec]

Posit

ion

[m]

Estimate fC

Estimated fCExpected fC

0 20 40 600

0.005

0.01

0.015

0.02

0.025

Time [sec]

Posit

ion

[m]

Estimate fv

Estimated fvExpected fv

Figure 2-6: Convergence of the adaptive parameters fC and fv

available, while the settling time is considerably elevated (around 100 seconds). Test

1 will be used as reference for comparison with the nonlinear controllers to be devel-

oped in this chapter.

Its pronounced slope shows the tipical profile of perturbance rejection, in this case the

friction on the table. Test 2 shows hard effects in friction and despite the fact that

it reaches the termination conditions (x < 0.01 [m/s] and e < 0.01 [m]) it cannot be

trusted as a solid reference of the performance of this controller because the friction

effects are highly evident in the uneven shape of the response of the controller. Fric-

tion is clearly evident in test 2 between 20 and 60 seconds. That slow slope represents

static friction, which means that the satellite actually was stopped by friction.

36

0 50 100 1500

0.1

0.2

0.3

0.4

0.5

0.6

Time [s]

Posit

ion

[m]

Test1Test2

Figure 2-7: Dynamic performance of the current PID controller, xd = 0.5 [m]

Figure 2-8 shows the tuning of the proposed controller. In the present design three

dynamic performances are higly desired: fast rise time, high damping and no over-

shoot. The requirement of no overshoot lays in the fact that the present work is

applied in Rendezvous and Docking (RDV) systems. Even a slight overshoot will

cause an unnecessary collision.

Increasing the value of γ does not affect in a visible way the desired dynamic perfor-

mances. It rather helps to a faster convergence of the estimated values which in turn

leads to a better tracking performance. After a series of hardware experimentations,

it has been proven that for this particular controller a high value of gamma close or

higher than 10 leads to unstability of the system. The control gain k gain acts as a

proportional gain and this effect can be seen because the increasing of k reduces the

rise time while increasing the damping in the system.

37

0 10 20 30 40 50 60 70−0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

Time [s]

Posi

tion

[m]

gamma=0.7 k−0.7gamma=1 k=1gamma=1 k=2gamma=2 k=5

Figure 2-8: Tuning of the Nonlinear Controller

After extensive testing, it has been determined that the values for the adaptation

rate γ and the control gain k should be 2 and 5 respectivelly. This allows for a quick

convergence of the estimated parameters and a good response in the case of sliding

friction.

In real-world applications, overshoot happens and it is taken into account by adding

impact suppresion devices. In the present case it will be higly desired to reduce this

overshoot within the centimeter level and reach a millimetric accuracy.

Figure 2-9 compares the dynamical performances of the PID controller versus the

proposed Nonlinear Adaptive Controller. While the rising time is the same, the pro-

posed controller has a higher damping and does not overshoot. and the damping are

pretty much the same, the fact is that the present controller does not overshoot at

all and reaches the target position in a time by far inferior to the PID controller.

38

0 50 100 150−0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Time [s]

Posi

tion

[m]

NonlinearPID

Figure 2-9: Comparison of the dynamical performances of the proposed controller vs.the actual controller

2.5.3 Summary

The present chapter introduces the basic concepts of nonlinear adaptive control ap-

plied to the case of the double integrator system for the tracking problem under linear

parametrizable friction with unknown coefficients. This development will be useful

both for the development of the present thesis work and the future activities of the

SSL. In the concrete case of consistent docking, some of the higher level maneuvers

such as CAM or Path Planning have to be tested in ground in order to fully optimize

the resources in orbit. This controller certainly allows better tracking without over-

shoot, which is vital for the current docking applications at the SSL.

39

40

Chapter 3

Robust H∞ Control

3.1 Introduction

In this chapter a suboptimalH∞ robust feedback controller is developed to solve single

and multiple body trajectory tracking problems. While the ISS normally provides

the satellites a friendly environment against external perturbations and even if the

dynamic behavior of the system might be well known there exist uncertainty in the

behavior of the real system because of neglegted dynamics or physical behavior of the

subsystems.

3.2 Satellite Dynamics

Two assumptions are taken to setup the problem statement:

• H1: The satellites are considered as a single mass point.

• H2: Position and attitude dynamics are decoupled.

The equations of motion are reduced to:

m x = F (3.1)

I θ = T (3.2)

41

Where x represents the position vector x = [x y z] and F = [Fx Fy Fz] represents the

thrust vector in the axial directions. I represents the Inertia tensor and θ = [α θ ψ]

represents the angular position vector expressed in the pitch, roll and yaw angles.

T=[Tx Ty Tz] is the torque vector used in the rotational dynamics. The scope of this

work corresponds to the translational dynamics given that the rotational dynamics

are highly nonlinear and the current attitude control algorithms developed in the

SPHERES program have been found to have a good performance. Equation 3.1 can

be represented in the Laplace space L by the transfer function:

G(s) =1

s2(3.3)

This system is uncontrollable because the transfer function has two poles in the jω

axis. It is necessary to slightly perturb each of the elements of the plant as follows:

G(s) =1

(s+ ε)2(3.4)

where ε = 0.0001 is a small perturbation required to guarantee the controlability of

the system. This system can be represented as the following continuous time state-

space system:

x = Ax + BF (3.5)

y = Cx + DF (3.6)

where the minimal realization of the state matrices can be expressed as:

A =

-2x10−5 -1x10−6

0 1

B =

0

1

C =[

0 0.2326

]D = 0

(3.7)

The use of small perturbations in a second order plant is found in the literature and

does not affect the behaviour of the system [11, pp 19-321].

42

3.3 H∞ Control

3.3.1 The H∞ standard problem

In the H∞ problem the synthesis scheme considered is shown in Figure 3-1, The

vector u represents the control inputs while the vector y represents the available

measurements. The vector w reunites the considered exterior inputs (i.e. reference

signals, disturbances, noises) which can be the input of the shaper filters to be chosen.

Finally, vector z includes the signals chosen to characterize the good functioning of

the feedback cntrol system, which are generally obtained from the signals existing in

the feedback control loop [[9] pp.480]

Figure 3-1: Standard H∞ Setup

The augmented plant P includes the system to be controlled and the shaping fil-

ters of the performance outputs. This plant can be represented by the equations:

x = Ax + B1 w + B2 u (3.8)

z = C1 x + D11 w + D21 u (3.9)

y = C2 x + D21 w + D12 u (3.10)

Where D12 is zero because the present work does not take into account the delays in

the system.

43

The augmented plant P can be also represented by its compact matrix representation:

P =

A B1 B2

C1 D11 0

C2 D21 D22

(3.11)

The optimal H∞ problem is to find all admissible controllers K(s) which minimize

the H∞ norm of the feedback system, i.e. all admissible controllers that minimize

‖P (s)‖∞. The norm ‖‖∞ can be physically seen as a measurement of the energy

performance of the system, i.e. the energy gain from input u to the output y ([7])

and is described by:

‖P (s)‖∞ = supw∈R

√λ(P (jw)P (−jw)T (3.12)

Finding an optimal controller is difficult and the optimal controllers for the H∞ prob-

lem are not unique. They can be viewed as the limit case for suoptimal controllers

and are not explicitly constructed. In this section, the suboptimal H∞ problem is

considered. The suboptimal H∞ controller K(s) is found for a given γ > 0 when

‖P (s)∞‖ < γ. The optimization problem consist in finding a value γopt such that, for

any admissible controller:

γopt = inf {‖P (s)| ∞ } (3.13)

For γ = γopt there are no suboptimal controllers. A controller is said to be admissible

if it stabilises internally the system P (s) and it is proper, i.e. the vector D in the

system 3.1 is zero.

3.3.2 Characterization Theorem for Output Feedback

The H∞ suboptimal controllers assume the following properties:

44

(A1) (A,B1) is stabilizable and (C1,A) is detectable.

(A2) (A,B2) is stabilizable and (C2,A) is detectable.

(A3) DT12 C1=0 and DT

12 D12=I

(A4) B1 DT21=0 and D21 DT

21=I

(A5) D11=0 and D22=0

It can be seen ([11], pp. 165-170) that the previous asumptions are sufficient condition

for the stability of P (s) and the internal stability of the closed-loop system. In fact

(A1), (A3) and (A4) imply that the feedback loop is internally stable if and only if

G(s) ∈ H∞. The existence of the H∞ controllers via Riccati equations can be easily

found in the classical textbooks such as [11], [20] and [23] and therefore it is ommited

here.

3.3.3 Mixed Sensitivity Design

The proposed methodology is to pose the H∞ standard problem using a mixed sensi-

tivity desing. For doing so, the original plan G is mathematically extended by using

shaping weights that contain the design goals of the system. These weights usually

have the form of a nth order filter. The extended plant is denoted by P.

The sensitivity function is a semi-proper transfer function with an infinite H2 norm.

This function represents the closed loop transfer function from the output distur-

bances to the outputs. The proposed controller in Fig. 3-2 is an one degree of

freedom (DOF) controller which can be shapen to have better dynamic performances

or perturbance rejection characteristics but not both at the same time. The following

terminology is introduced ([20]):

L = GK (3.14)

S = (I + GK)−1 = (I + L)−1 (3.15)

T = (I + GK)−1GK = (I + L)−1L (3.16)

45

Figure 3-2: Block Diagram for the one DOF controller

Where L is the loop transfer function, S is the sensitivity function and T is the com-

plementary sensitivity function. G is the plant to be controlled and K is the feedback

controller. The term sensitivity function is natural because S gives the sensitivity re-

duction afforded by feedback. The term complementary sensitivity follows from the

identity:

S + T = I (3.17)

Where I is the identity matrix. In the present work the approach taken is to shape

the controller by using the sensitivity function S. This approach is known as the

KS/S mixed-sensitivity optimization. In this approach S and KS have to be shaped,

resulting in the augmented plant P:

P =

W1 W1G

0 −W2

I G

(3.18)

where W1 and W2 are the weightings of the noise inputs. It can be shown [[20],pp.

188] that the closed-loop transfer function results in:

Fl (P,K) =

W1S

W2KS

(3.19)

46

In the present work, the control input filter W2 has been chosen to be equal to

1, or I in the multivariable case. This helps to simplify the model. A common filter-

ing suggested by [2] and [20] for the performance output filter W1 is:

Figure 3-3: S/KS sensitivity optimization

W1 =S/M + ωBs+ ωB ∗ A

(3.20)

where A << 1 = 0.0001, M ≥ 1 = 1.5 and ωB = 0.483 [Hz] is the cutoff frequency of

the system. Figure 3-4 shows the frequency response of the weighting W1. The func-

tion 1/W1 is considered as the upper bound of S. It is equal to A at low frequencies

and equal to M at high frequencies, and the asymptote crosses 1 at the frequency wb,

which is usually the bandwidth requierement.

47

10−5

10−4

10−3

10−2

10−1

100

101

102

−10

0

10

20

30

40

50

60M

agni

tude

(dB)

Bode Diagram

Figure 3-4: Magnitude Plot of W1

3.4 Results

3.4.1 Simulation Results

The experimental methodology chosen is similar to the one applied in the previous

chapter. The satellite will move to the position xd = 0.3 [m] and yd = 0.3 [m] at zero

velocity with a PID controller. Once this start position is reached the controller is

tested by switching controllers and having as its new target xd = 0.8

One of the objectives in the synthesis of a H∞ controller is to reach γopt = 1. This

is not a crucial objective, rather it is a design goal. The synthesis of the controller

hereby developed reached a value of γ = 1.4801. This can be seen in Figure 3-5. In

a ideal controller the sensitivity function S is expected to not surpass the function

1/W1. Once again, this characteristic is not critical and the values of the weights

were selected in function of the performance of the controller in the discrete-time

SPHERES dynamics simulator.

Figure 3-6 shows the simulation results of the controller for a step input response of

0.5[m] . The controller was implemented at Ts = 2[s]. The controller K(s) is trans-

formed from its state space form into a discrete filter. The H∞ controller has a faster

rise time than the PID controller. The LQR controller has the best damping and rise

time among the three controllers.

48

−1 0 1 2

−0.6

−0.5

−0.4

−0.3

−0.2

−0.1

0

0.1

Freq (rad/sec)

Mag

nitu

de

S1/W

s

WsS

Figure 3-5: Weighted Sensitivity Test

Two important aspects have to be adressed: The first is that none of these con-

trollers overshoot, which is the ideal case and seems logical trough simulation. This

is highly required for the RDV procedures where a significant overshoot might lead

towards collision.

The second is that the LQR is normally expected to have a better dynamic per-

formance with respect to the H∞ controller. This is because the LQR controller is an

optimal H2 controller especifically designed to have a high performance while the H∞controller is designed to compensate the uncertainties in the system. A H∞ controller

has an internal performance controller plus a filter. Thus, the H∞ is not expected to

better those dynamical performances.

In the one DOF H∞ problem the robustness of the system and the performance

are antagonist by definition. The performance of the proposed controller could be

improved with a good robustness by modifying the problem into a two DOF robust

problem. This case will not be adressed in the present work. The controller imple-

49

0 10 20 30 40 50 600

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Time [sec]

Posit

ion

[m/s

]

Tracking Response

HinfPIDLQR

Figure 3-6: Performance Comparison. Step Input of 0.5[m]

mentation is described in algorithm 3. In the present case the controller has been

implemented as a finite impulsional response:

u [k] = b0 x [k] + b1 x [k− 1]; (3.21)

where b0 and b1 can be seen as gains of the filter. By doing so, the implementation is

easy and the computational load is the same of a PID or LQR controller. The variable

istate represents a internal state used by the controller. It contains the previous value

of the state x. T is the control period. The value of istate(2) is implemented as a first

order recurrence equation to take into account the velocity. This will improve the

performance of the proposed controller.

Another way to implement this controller is by using the matricial form of the con-

troller. This will result in a discrete-time state space representation with the matrices

50

Alg. 3 ctrl pos hinf(e, istate). A robust linear control algorithm to track a determinedtarget.Input: e, istateOutput: u

1 begin2 define T , b0, b1;3 u = istate(1) = e;4 u = b0 ∗ istate(1) − b1 ∗ istate(2);5 u = istate(2) = e+ e ∗ T ;6 end

Ac, Bc and Cc obtained from the synthesis of the controller and by using the error

as the input. They are the space state obtained from the controller. Another useful

form of representation is the use of a recurrence equation derived from the discrete

transfer function of the controller. Both algorithms require more computation than

the proposed method.

3.4.2 Experimental Results

Figure 3-7 shows the performance of the proposed controller under nominal conditions

in the table. A controller was synthetised for the flat floor table. This controller was

complemented with a integration error algorithm to improve its performance. It can

be easily seen that the proposed controller sligthly overshoot in the first synthesis

and performs underdamping in the second synthesis. These controllers have a value

of γ = 1.4 (Test 1) and γ = 1.30 (Test 2) with a cuttoff frequency wb = 0.735 and

wb = 0.5. The state matrix C of the system 3.6 has been reemplaced by C = [0 0.0747]

because the test weight is 13.5 [Kg].

Both controllers show a better dynamic performance than the PID. In this case the

dynamic performance was improved for different gains in the error integration algo-

rithm. Previous synthesis were developed in the discrete-time simulator with critical

dampening. However the friction nature of the table and most important the slope

51

0 20 40 60 80 100 1200

0.1

0.2

0.3

0.4

0.5

0.6

PIDHinf (Test1)Hinf (Test2)

Figure 3-7: Performance Comparison. Step Input of 0.5[m]

of the table are not taken into account in that simulator and the resulting controller

does not performs as expected. The second synthesis leads to a filter of the form:

u [k] = 1.071 [k] + 0.4953 [k− 1] + 0.3 ∗ eint (3.22)

3.5 Summary

The present Chapter introduced a basic robust H∞ controller. The synthesis of the

present controller was based in the assumptions of a 6DOF work environment and

no coupling between the attitude controller and the position controller. Previous ex-

perience aboard the ISS show that both assumptions are true. In ground testing the

evidence of a coupling between both controllers given the particular geometry of the

air carriage is evident.

52

A modified H∞ controller for the air table is described. For this particular case

an error integration algorithm was added to the proposed controller in order to reach

the final position. This controller was developed as a way to demonstrate the imple-

mentation of the H∞ controller proposed in this work. Only flight experience will

show whether the H∞ controller has accomplished its objectives.

53

54

Chapter 4

Consistent Rendezvous and

Docking

4.1 Introduction

Rendezvous and docking or berthing (RVD /B) is a key operational technology, re-

quired for many missions involving more than one spacecraft. Control techniques have

been successfully implemented in the automated rendezvous and docking of space-

crafts requiring some degree of manual control. However, when the uncertainty of the

system increases the performance of RVD /B may diminish to unacceptable levels

with potential mission failures. The testing of new technologies such as the AVGS

navigation proposed in the DART mission [19] are too expensive to fail.

4.1.1 Consistent Docking

The idea of Consistent Docking comes to answer the problem of RVD aiming towards

fully autonomy of the process. This includes the integration of several algorithms pre-

viously developed or in development at the SSL such as path planning, fault detection,

plume imprigment, collision avoidance, position and attitude controllers, among oth-

ers. The key contribution of the present work is the addition of the adaptive nonlinear

and robust linear controllers with application in grond and space respectivelly.

55

The complete automation of the rendevouz procedures are crucial in the next phases

of space exploration. As menctioned in the introduction of the present work, there will

not be real-time manned surveillance capability for any RVD maneuver executed in a

planetary mission because of the delays in communication. Besides, this technology

can be applied for the servicing of satellites and space probes without the intervention

of astronauts. With the retiring of the Space Shuttle any manned satellite servicing

other than the Hubble servicing mission is discarded in the near future.

The criteria of success expected in this technology for purpose research at the SSL

is to have a safe automated docking with a 90 percent of success as a starting point,

despite the evolving nature of the space environment. The ISS allows to develop the

required GNC algorithms iteratively in a relevant environment and with a certain

safety in case of failure.

Two applications are foreseen in the framework of the present work. The first is

the development of docking procedures for the ground testing of the SWARM mod-

ular spacecraft concept. This will be explained in the next section. The second

application is the testing of docking procedures aboard the ISS and the study of the

response of these docking procedures under a unexpected thruster failure.

4.2 Orbital Assembly

One of the concepts being curently explored at the SSL is the Self-assembling, Wire-

less, Autonomous, Reconfigurable Modules (SWARM). This is an innovative approach

to modular fabrication and in-space robotic assembly of large scale systems. The

SWARM concept uses formation flown spacecraft, containing multiple universal dock-

ing ports (Figure 4-2) whose serve to dock with modular elements and maneuver them

to dock with other similar elements.

56

In the process, systems can be assembled that are much larger than what can be

fit or folded into a launch vehicle fairing, or what can be launched on a single vehicle.

Furthermore, such modularity will allow jettison of failed components, upgrade of

obsolete technology, and amortization of design costs across multiple missions.

The fabrication of modular components yields savings associated with large produc-

Figure 4-1: SWARM modules being tested at the MSCF flat flor

tion volume and automated integration and test. In-space assembly permits staged

deployment on an affordable basis with potential cost and risk reduction by using

smaller, less-complex launch vehicles. It also decouples stowed launch geometry from

deployed operational geometry. The SWARM approach to developing such an archi-

tecture consists of developing:

• Universal docking ports with wireless command and data handling (CDH)

• Space tug maneuver and docking control

• Modular components with multiple docking ports

During the first phase of the SWARM project a methodology for integrating these

technologies was successfully developed. Basic SWARM functionality for a particular

mission scenario (Telescope Assembly) was succesfully demonstrated on the MSFC

flat-floor (Figure 4-1). This included docking and reconfiguration of three modules

representative of a modular large space structure using elements of the ARGOS tele-

scope and demonstrated the feasibility of the SWARM approach.

57

Figure 4-2: SWARM Docking Port

4.3 Proposed Docking Algorithm

This section discuses the docking algorithm to be used in the present test. It has been

succesfully tested aboard the ISS and it is the basis of future testing for docking oper-

ations at the SSL both in ground facilities and aboard the ISS. This algorithm uses a

path planning algorithm to generate intermediate trajectories which are to be follown

by the satellite. The current PID controller has acceptable tracking performances but

in past docking experiments aboard the ISS the overshoot leaded to collision of the

satellites. The docking methodology follown is depicted in figure 4-3.

The use of the proposed controllers and the path planning algorithm allows the chaser

satellite to get infront of the docking port situated in the front of the target satellite.

The use of the path planner helps to eliminate assumptions of an initial relative con-

figuration between the satellites. This maneuver is divided in five phases:

58

Figure 4-3: Docking to a Fixed target Satellite Facing Backwards

1. Planning: The chaser satellite generates a safe trajectory with the path planner

in order to bring the satellite to an alignment position at a safe distance with

both docking ports (DP) facing each other while avoiding any possible collision

between both satellites. This step takes 10 milliseconds.

2. DP Axis Alignment: The chaser satellite follows the path with the position

and attitude controllers. In the present work the path includes a waypoint 40

seconds into the maneuver where the path is replanned. This helps to improve

the precision of the satellite when arriving to the DP alignment point.

3. Close In: The satellite replans a short trajectory to reach a closer point to the

satellite.

4. Berthing: Before berthing or docking, the satellite has to fulfill a number of

tight constraints in attitude and position. The control tries to bring the satellite

whitin these specifications. Considering that the satellite might not be able to

reach that precision, there is a timeout in the berthing maneuver. This is based

in the assumption that the satellite might be close enough to the target point

but out of the restrictive boundary set by the specifications.

59

5. Capture: The chaser satellite moves towards its target and both docking port

enter into contact.

Given the restrictive nature of the berthing conditions there is a good possibility that

the maneuver fails. The SPHERES satellites aboard the ISS use velcro as a safe,

inexpensive solution that allows to explore docking maneuvers. Some problems arise,

such as plume impingement and hard contact with the chaser. These problems will

not be adressed in the framework of the present thesis.

In the case of ground procedures the problem becomes harder because of surface

imperfections. Even considering an adaptive controller the surface might have irregu-

larities that can easily lead the testing hardware in the wrong direction. At the time

of the writing of this thesis the 6 DOF test aboard the ISS haven’t been run because

of factors out of the author’s control. The following results will focus extensivelly on

the ground work rather than the ISS research which hopefully will be performend in

the early fall.

4.4 Experimental Results

The results described in this section include the air table at the SSL laboratory and

the flat floor. The flat floor facility provides researchers with a bigger surface for

experiments, in comparison with the limited surface on the air table (about 1.5 [m2]).

The proposed experimental methodology for the automated docking maneuver is the

following:

• Both satellites are put in front of each other, ibetween 70-100 centimeters. Both

docking ports are facing each other.

• The chaser satellite will adjust its attitude to align both docking ports. In this

test the target satellite has been set to keep position.

• The chaser satellite follows the path, reaches the waypoint and replans the new

path.

60

• After following the close-in path the satellite reaches the berthing position. It

stays there until the position error is less than 5 millimeters or it reaches its

timeout after 60 seconds in this maneuver.

• Finally, once the conditions are fullfilled or a timeout happens the chaser satel-

lite moves towards the target satellite. If succesfully executed, this maneuver

should dock both satellites within a tolerance of 1 cm.

4.4.1 Docking Results in the Air Table

Figure 4-4 shows the tracking capabilities of one test of the adaptive controller. It

can be seen that globally the controller does act as a path follower. However there is

a steady tracking error of about 0.8 cm. This can be explained by the fact that the

satellite was overcoming a slope in the Y direction due to a slight inclination in the

table.

0.9 1 1.1 1.2 1.3 1.4 1.5 1.60.4

0.45

0.5

0.55

0.6

0.65Docking Test in the Flat Floor. Nonlinear Adaptive Controller

x [m]

y [m

]

Real PositionTarget Position

Figure 4-4: Docking Test with the Adaptive Controller at the Air Table

61

The nonlinear controller does not handle well the slope at this point. The black dots

show the waypoints generated by the path planner. The controller reach the way-

points and performs well the final approach. Even given that the controller does not

meet the berthing conditions, it reach the docking point within the tolerances. In the

experiment shown above the position error at the point of contact was 0.64 [cm]. This

allows a docking whitin the tolerance. This tolerance appears from the specific design

of the docking port shown previously in figure 4-2. The hole for the pin is 2.5 cm wide.

Being withing 1 [cm] will allow the satellite to be well inside the chanferred area

of the hole, so even if the internal hole is just 1 [cm] in diameter, the chmamfered

hole will contain the pin. Harware implementation of the docking port will be left as

future work because the path planner requieres to be reconfigured. At this point the

path planner considers its default docking face the -X plane of the satellite, which

contains the velcro face. Future work will include adding the current version of the

docking port which includes its own metrology system that can be used for close range

tracking. By doing so, the precision at the contact point and the contact velocity can

be handled more precisely than with the global metrology used throughout this thesis.

Another explanation of the tracking error can be the non-optimality of the gains

used. Optimal gains specified for the tracking problem will lead to a better tracking

precision and reach safely the docking port. Finally, the lack of accuracy in the fric-

tion model has to be taken into account. The Coloumb-Viscous model does not take

into account pre-sliding deplacement, which can be unexpectedly present across the

table in the form of a scratch or a surface irregularity. This can be handle with the

internal state models, whose take that part into account but are strictly nonlinearly

parametrizable.

62

4.5 Docking Results on the Flat Floor

4.5.1 Docking Test with the PID Controller

The use of the flat floor provides scientist the opportunity of working with a broader

range of experiments that were not viable in the reduced surface of the air carriage.

However the friction characteristics are not the same. As a matter of fact, the control

of the satellite with the air carriage in the flat floor is a more complex task given that

the uneveness of the surface is more important. Figure 4-5 shows the response of the

current PID controller in a docking experiment.

0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.650.04

0.05

0.06

0.07

0.08

0.09

0.1

0.11

0.12

0.13

0.14Docking Test in the Flat Floor. PID Controller

x [m]

y [m

]

Real PositionTarget Position

Figure 4-5: Docking Test with the PID Controller at the Flat Floor

The tracking is good overall but the control activity is really demanding for the satel-

lite. This means a huge use of resources such as CO2 for the propulsion and batteries

for the soneloids in the nozzels. The controller starts to diverge before reaching the

first waypoint.

63

The path shape can be explained by a reconfiguration made by the path planner

because of a existent timeout of 60 seconds between the start of the test and the way-

point.the controller still tries to track the new path and reaches the berthing position.

However, while the controller is in the berthing position integrates the error and this

results in a completelly wrong contact point. In the experiment shown above the PID

controller misses the target by more than 5 centimeters.

4.5.2 Docking Test with the Adaptive Controller

Figure 4-6 shows some experiments run in the finding of the new gains for the adap-

tive controller. It can be seen that the PID controller has a static error of around 9

cm.

0 20 40 60 80 100 1200

0.1

0.2

0.3

0.4

0.5

0.6

Time [s]

Posit

ion

[m]

Step Input Test in the Flat Floor with xd=0.5 [m]

PIDNonlinear (Test1)Nonlinear (Test2)

Figure 4-6: Step Input response of the Flat Floor Controller xd=0.5 [m] at the FlatFloor

64

The particular shape of the control response in this case responds to the fact that

the satellite reached an uneven surface and it was unable to overcome, the thrust

level required sent to the satellite by the PID controller do not overcome the friction

effects. The flat surface in the response plot means that the controller is overcoming

the friction slowly.

The first experiment with the nonlinear controller (Test1, in figure 4-6 shows that

by using h the controller gain k=10 and the adaptation rate γ=3 has a slow rising

time and does reach the target position at the end, having a slight overshoot of two

centimenters.The second experiment (Test2) shows that a synthesis with k=20 and

γ=3 and it overshoots slightly but does reach the target position.

Nevertheless a more precise model of the friction has to be taken into account given

the fact that the presliding friction effects are not taken into account in the curent

model. This leads to unexpected performances of the controller in some tracking test.

It has to be stated that both the adaptation rate and the control gain have not been

found by numerical optimization. This could help to reach an optimal performance for

tracking problems. This optimization is not treated in the framework of the present

work. These values are found trough direct experimentation and their performance

has been evaluated in considering both tracking error, steady state error, overshoot

and rise time.

Figure 4-7 shows the docking results of the adaptive controller at the Flat Floor

using γ=3 and k=10. This controller does not perform very well in the tracking

performance. At the time of the writing of this work the exact values for parameters

of the adaptive controller at the Flat Floor haven’t been found. It has to be stated

that the scale of this plot shows a big error on the tracking. However the final error

is actually 2.2 [cm].

65

0.7 0.8 0.9 1 1.1 1.2 1.3 1.40.63

0.64

0.65

0.66

0.67

0.68

0.69

0.7

0.71

x [m]

y [m

]

Docking Test in the Flat Floor. Nonlinear Adaptive Controller

Real PositionTarget Position

Figure 4-7: Docking test of the Adaptive Controller at the Flat Floor

Once again, the path planner recalculates the path when the timeout has been reached

and the controller tries unsuccesfully to track this path. This error might be also

caused by the omission of other friction dynamics that can be included in nonlinear

parametrizable models with internal state such as LuGre or Maxwell Generalized.

4.5.3 Docking Test with the H∞ Controller

Some experiments were run by using the H∞ controller. A H∞ controller was iden-

tified with the However, given the fact that the nonlinearities are not taken into

account in the synthesis of the controller, it cannot be expected that the controller

will perform correctly. This controller is able to reject external perturbations in a

6DOF environment. During ground testing the coupling between the position and

the attitude controller is strongly evident and the assumptions taken in Chapter 3

are not true anymore.

66

Figure 4-8 shows the preliminary results of a docking test with the H∞ controller

modified for operations on the air table. A controller for testing on the flat floor will

not be adressed in the present work. It can be seen that the controller does follow

the track but the cumulated tracking error originated from the omision of the friction

dynamics in the synthesis of the controller lead to a final error of more than 4 [cm].

−0.2 −0.1 0 0.1 0.2 0.3 0.4 0.50.5

0.52

0.54

0.56

0.58

0.6

0.62

0.64

x [m]

y [m

]

Docking Test in the Flat Floor. Hinf Controller

Real PositionTarget Position

Figure 4-8: Docking Test with the H∞ controller. Tracking of the Trajectory

At the time of the writing of this thesis the H∞ controller step input test and the

docking test with theH∞ controller were not performed aboard the ISS. These test are

scheduled in four experiments in Test Session 13, scheduled to be flown in September

2008. The experiments include the comparison of docking capabilites of the proposed

controller under thruster malfunction as a perturbance of the system. This test is

compared with another experiment in the same Session by using the LQR controller

developed and to be tested in Test Session 12, at the end of August.

67

The other test to be flown aboard the ISS as a part of the present work include

also a comparison of the tracking performance of the proposed controller with respect

of the PID and the LQR controller in the same maneuver. The maneuver chosen is

the path tracking of a Spiral maneuver. Figure ?? shows the tracking performance

of the proposed test based on simulation. It can be seen that the LQR is the best

tracking controller with a final error of 6 cm. The PID controller has a final error of

12 cm and the H∞ controller has a final error of 10 cm. This was expected because

the H∞ controller is not specified for performance and it is expected to have a infe-

rior performance with respect of the H∞ controller. In the other hand, the proposed

controller has a better performance if compared with the PID controller and this is

only for tracking problems.

4.5.4 Summary

The present chapter explored diverse experiments with the docking maneuvers by

using different controllers. The test were run by using the PID, Nonlinear Adaptive

and H∞ controllers. Given that these test were run in a frictioning workspace it was

the logical result that the nonlinear adaptive controller is evidently the best solution.

The PID controller has been extensively used in the past because of the madurity of

that control. It is expected that with time the nonlinear adaptive controller becomes

the standard controller for ground testing at the SSL.

In overall the adaptive controller does well its work of friction compensation. However

it has been seen that tracking performance is an issue to be adressed. Mathematical

optimization will bring important feedback to reach an optimal controller with exce-

lent tracking capabilities.

The case of perturbance rejection has not been adressed in this work. Test Ses-

sion 13 will bring invaluable data in the proposed H∞ test performed in a relevant

environment. This leads to a incremental work for Test Session 14 where other robust

linear controllers will be developed aiming towards decentralized control.

68

Chapter 5

Conclusions

5.0.5 Thesis Summary

The present work has shown the development of robust nonlinear and linear con-

trollers and their application to flight formation task. These controllers can be used

in any formation flight, docking or multi body spacecraft maneuvering for the ground

test and space flight cases respectivelly. The objective of the present thesis was to

strongly pursue the use of the proposed controllers towards a fully consisted auto-

mated rendezvous and docking of the satellites. The development of the proposed

controllers is another increasing task of many others to be developed in order to full-

fill a real consisted RVD algorithm tested under extreme operational conditions in a

relevant environment.

Chapter 2 proposes an adaptive nonlinear controller to compensate the friction during

ground testing. The metrics evaluated to measure the improvement of this controller

with respect to the previous PID controller were no overshoot, no static error and a

fast rise time without incurring in overshoot. This controller has been proven to be

succesful in basis of the metrics imposed and considering the step input test. The

step input test is a standard test in control analysis and design. However this control

should be modified for specific applications. In particular, the tracking performance

of the proposed controller has to be improved.

69

Overshoot happens in the controllers on the flat floor. This is required because the

gains had been increased to overcome a friction that experimentally has been proven

to be more challenging that the work at the air table. Even with the optimal gains

found trough mathematical optimization, there is a fair amount of work to be done

in order to At the time of the writing of this thesis his controller has been succesfully

used in applications other than docking. However tracking must be improved given

the fact that tracking error is a metric to be considered in several other applications

such as star navigation or interferometry.

In Chapter 3 H∞ control is introduced as an approach of how to deal with uncer-

tainties. The deveopment of the present controller is a first step in the way of more

complex robust algorithms. Robust lineal control is a topic not well explored in the

current level of the SPHERES program. The ATV mission shown that perturbance

rejection in LEO and in approach/berthing/docking can be done in a more efficient

way that other controllers such as Eigenvalue Assignment and PID [19]. In the case

of a robust lineal controller for spaceflight in a outer space environment, mathemat-

ical computation might increase and the dynamics of the perturbation have to be

linearized. Other solution is to use a complex LTV lineal synthesis. This work is yet

to be done at least for the case of fully automated docking procedures

H∞ is a technique not often used in spaceflight because of the risk that it might

imply to use a new technology and because of the madurity of the present techniques.

With the computational power of onboard calculators increasing and a better under-

standing of the space environment, the H∞ applications in space flight environment

will increase with time.

The robust controller has not been tested aboard the ISS by the time of the writting

of the present thesis. The present year has seen an important activity in the construc-

tion of the orbital outpost. Test Session 12 has been postponed several times and

it will bring invaluable data from docking test with LQR algorithms during August

70

30th, 2008. TTest Session 13, to be flown not earlier than September 6th, 2008, will

validate the same algorithm under thruster malfunction and using the proposed H∞Controller. Also, the tracking performance of the proposed controller will be tested

by using a three-satellites spiral algorithm where each of the satellites has a different

controlller.

While the H∞ accounts for the parametric uncertainty of unmodeled dynamics, the

LQR is better performant. A tradeoff must be found in the selected application. The

use of a 2DOF H∞ controller must lead to better dynamical performances while in-

creasing the robustness of the system. The selected metric for the H∞ test is rejection

perturbance. So far simulations show a better rejection perturbance than the PID

controller. However this simulation does not take into account all the dynamics and

therefore is not conclusive. Flight data will pave the road for the improvement of the

proposed H∞ controller.

Chapter 4 shows the use of both controllers to the docking maneuvers. The pro-

posed docking architecture shown in Chapter 4 has been succesfully used at the ISS

beign the first docking experiment with a path planner used on orbit. The path plan-

ner has proven to be a helpful resource for this architecture. However, as it has been

seen in previous experiments, the low level controllers must be improved to better

the tracking capabilities and avoid overshooting. The use of LQR controllers aboard

the ISS for docking purposes will bring a insight on which performances are required

for the docking problem.

In the case of ground operations the nonlinear adaptive controller has ben tested

extensively in the air table. The work in the flat floor is fairly new. Tests in the air

table are most of the time succesfuls but still there is a fair amount of work to do

if a more complex dynamica model has to be included (multi body problem) or if

a new friction model is to be considered. In that case, the present synthesis is not

valable anymore and a new nonlinear adaptive controller has to be designed unless

71

the nonlinearities are linearized The docking algorithms H∞ controller will be tested

aboad the ISS in the early fall 2008. Simulation has shown the feasibility of this con-

troller and the potential success of the mission to be flown. However, as it happens

more often than aerospace enginners would like, errors can happen. Therefore only

until the results come from the ISS it can be stated that the developed controller is

succesful and has fullfilled its objectives

5.1 Contributions

The most important contributions of the thesis work hereby present are related with

the area of robust control. Both nonlinear controllers and robust linear controllers

have been developed in the framework of this thesis. One of the key points in the

work hereby presented is a high performance nonlinear controller which will

be the standard position controller during ground testing at the SSL not

only for the SPHERES projects and for future aplications.

The robust controller hereby proposed is expected to be the cornerstone of a con-

sistent docking algorithm. This technology is in development and it will take a long

time to have a full in-space tested version of this consistent docking algorithm The

robust controller is expected to assure the robustness of the system despite

the neglected satellite dynamics given the simplifications taken in the synthesis

of the proposed controller. Besides, it is highly difficult to have an accurate model

of the whole system with uncertainties derived from continuos-time dynamics of the

actuators and discrete-time delays from communication and data processing.

Finally, the key contribution of the present work is the development of controllers

and the use of existent docking algorithms to prove the ability to conduct a con-

sistent docking in ground with a success rate superior than 90 percent .

At this point this capability has been reached in the air table. Work in the optimal

This is key for the testing of the SWARM program to be done during the winter of

72

2008 at the Marshall Space Flight Center in Alabama.

5.2 Future Work

The nonlinear controller still can be improved. Some problems have been seen during

the pre-sliding displacement, this is self-correcting given the nature of the proposed

control algorithm. Other nonlinear parametrizable models such as LuGre or Elasto-

Plastic might help to solve this problem.

From the operational point of view, the proposed controller has fullfilled the expec-

tations. In the case of continuous breaking algorithms such as stop-and-stare control

algorithms proposed for satellite imagery, this might be a problem that will be need

to be adressed.

The nonlinear controller has proven to be an almost ideal controller for the proposed

task (milimeter-precision control with friction compensation). Experience shows that

this controller works fine under several surface discontinuities. Yet, there are other

dynamical characteristics that are to be improved, such as presliding displacement.

The present work shows that the proposed controller is far superior to the actual

controllers developed in the SSL for the ground testing. In the past, members of

the SPHERES team hesitated to fully verify their algorithms in the 2-DOF facility

because of the well known friction and gravitational effects.

Hopefully, the proposed controller will lead to a new culture of full testing before

the orbital testing and even more: to test new technology with hardware in develop-

ment, such as the case of the SWARM project.

It is clear that a lot of work remains left in the current docking technologies available

at the SSL. In comparison of the space agencies or the big companies, the solutions

proposed at the SSL must be viable technically and economically. Other docking

73

technologies might be suggested such as optical devices, visual-based navigation and

pattern recognition. Also. It has to be stated that the current facility at the ISS

has only the velcro faces as an emulation of this docking port. Ideally, in the future,

SPHERES might be upgraded with these docking ports already developed.

This would lead to a fully hardware-in-the-loop autonomous docking of spacecrafts,

which lead to a important contribution in the field by adding a series of maneuver-

ing conditions such as CAM maneuvers, obstacle avoidance, uncooperative docking,

thumbing docking and others that already have been adressed by other researchers at

the SSL but have to be proven with real docking hardware. It is important to recall

that the problem of docking is challenging for both hardware design and software

design. Both must be tested in order to reach a fully demonstrated docking ability.

Limitations in time did not allowed to have a full analysis of robustness and un-

certainties on the H∞ controller . Future work will be required on this subject to

take into account thruster saturation, uncertainties in the thrusters and noise from

the sensors.

74

Bibliography

[1] Mit space systems lab. http://ssl.mit.edu/about/. Last acces: August 18th,2008.

[2] I. Bates, D. Postlehwaite. Robust Multivariable Control of Aerospace Systems.Number 08 in Control and Simulation. DUP Science, 2002.

[3] N. Currie. Advancing technology and education for an human base on the moonand mars. Presentation to the Engineering Dean’s Conference, October 2006.

[4] P. Dupont, B. Armstrong, and .V Hayward. Elasto-plastic friction model: Con-tact compilance and stiction. In Proceedings of the American Control Conference,volume 2, pages 1072–1076, Chicago, IL, March 2000.

[5] W. Fehse. Automated Rendezvous and Docking of Spacecraft. CambridgeAerospace Series. Cambridge Univertsity Press, Cambridge, UK, 2003.

[6] M. Hilstad. A multi-vehicle testbed and interface framework for the developmentand verification of separated spacecraft control algorithms. Master’s thesis, Mas-sachusetts Institute of Technology, June 2005.

[7] J. How. Principles of optimal control. Lecture Notes, 2003.

[8] V. Lampaert, F. Al-Bender, and J. Swevers. A generealized maxwell-slip frictionmodel appropiate for control purposes. In Proceedings of the Physics and Con-trol Conference PHYSCON 2003, volume 1, pages 1170–1176, Moscow, Russia,August 2003.

[9] P.de Larminat, editor. Analysys and Control of Lineal Systems. 2003.

[10] W. Li. Adaptive Control for Robot Motion. PhD thesis, Massachusetts Instituteof Technology, May 1990.

[11] U. Mackenroth. Robust Control Systems: Theory and Case Studies. Springer-Verlag, 2004.

[12] C. Makkar, . W. E. Dixon, W.G. Sawyer, and G. Hu. A new continuouslydifferentiable friction model for control systems design. In Proceedings of the 2005IEEE/ASME International Conference on Advanced Intelligent Mechatronics,volume 1, pages 600–605, Monterey, CA, July 2005. IEEE/ASME.

75

[13] J. Mankins. Technology readiness levels. Technical report, Advanced ConceptsOffice, Office of Space Access and Technology, NASA, April 1995.

[14] S. Nolet. Development of a Guidance, Navigation and Control Architecture andValidation Process Enabling Autonomous Docking to a Tumbling Satellite. PhDthesis, Massachusetts Institute of Technology, June 2007.

[15] S. Nolet, E. Kong, and D.W. Miller. Autonomous docking algorithm developmentand experimentation. volume 5149. SPIE, 2004.

[16] S. Nolet, A. Saenz-Otero, D.W. Miller, and A. Fejzic. Spheres operations aboardthe iss: Maturation of gnc algorithms in microgravity. 2007.

[17] A. Saenz-Otero. The spheres satellite formation flight testbed: Design and initialcontrol. Master’s thesis, Massachusetts Institute of Technology, August 2000.

[18] A. Saenz-Otero. Design Principles for the Development of Space TechnologyMaturation Laboratories Aboard the International Space Station. PhD thesis,Massachusetts Institute of Technology, June 2005.

[19] N. Silva, M. Ganet, E. Goribus, and P. Augros. Autonomous rendezvous: Guid-ance and control in elliptical orbit. In AIIA Guidance, Navigation and ControlConference and Exibit, volume 1, Hilton Head, SC, August. AIAA.

[20] S. Skogestad and I. Postlethwaite. Multivariable Feedback Control: Analysis andDesign. Wiley, 1996.

[21] Li W. Slotine, J-J. Applied Nonlinear Control. Prentice Hall, 1991.

[22] S. Thiery, M. Kunze, A. Karimi, A. Curnier, and R. Longchamp. Friction mod-eling of a high precision positioning system. In Proceedings of the 2006 AmericanControl Conference, volume 1, pages 1863–1867, June 2006.

[23] K. Zhou, J. Doyle, and K. Glover. Robust and Optimal Control. Prentice Hall,1995.

76