Transcript

MASTER PROJECT Geodetic Engineering Laboratory

Integration of Skew-Redundant MEMS-IMU with GPS for Improved Navigation Performance

By

Stéphane Guerrier

Supervisors

Dr. Jan Skaloud & Adrian Wägli

June 2008

2

Table of Contents

1. INTRODUCTION ..................................................................................................................................... 11

1.1. CONTEXT OF THE STUDY...................................................................................................................... 11 1.2. OBJECTIVES......................................................................................................................................... 12 1.3. STRUCTURE OF THE RAPPORT.............................................................................................................. 13

2. THEORETICAL BASIS OF IMU REDUNDANCY................ .............................................................. 14

2.1. INTRODUCTION.................................................................................................................................... 14 2.2. GEOMETRICAL CONFIGURATIONS OF INERTIAL SENSORS....................................................................... 14

2.2.1. Information Filters ........................................................................................................................ 15 2.2.2. Partial Redundancy....................................................................................................................... 17

2.3. NOISE REDUCTION AND ESTIMATION.................................................................................................... 19 2.4. FAULT DETECTION AND ISOLATION......................................................................................................22

2.4.1. Introduction................................................................................................................................... 22 2.4.2. The Parity Space Method .............................................................................................................. 22

2.4.2.1. Measurement Model............................................................................................................................ 22 2.4.2.2. Parity Space Model............................................................................................................................. 23 2.4.2.3. Fault Detection Algorithm................................................................................................................... 24 2.4.2.4. Fault Isolation Algorithm.................................................................................................................... 28

2.5. INTEGRATION OF REDUNDANT IMUS WITH GPS................................................................................... 29 2.5.1. Single IMU Mechanization............................................................................................................ 29 2.5.2. Synthetic Mechanization ............................................................................................................... 29 2.5.3. Extended Mechanization ............................................................................................................... 32 2.5.4. Geometrically-Constrained Mechanization .................................................................................. 33 2.5.5. Algorithm Selection....................................................................................................................... 34

3. IMU ARCHITECTURES ......................................................................................................................... 35

3.1. INTRODUCTION.................................................................................................................................... 35 3.2. ASSESSMENT IN TWO DIMENSIONS........................................................................................................ 35 3.3. OPTIMAL CONFIGURATION WITH THREE SENSORS IN THREE DIMENSIONS.............................................. 36 3.4. OPTIMAL CONFIGURATION WITH FOUR SENSORS IN THREE DIMENSIONS............................................... 37 3.5. OPTIMAL CONFIGURATION WITH N SENSOR TRIADS............................................................................... 40 3.6. NOTE ON THE IMPACT OF SENSOR FAILURES......................................................................................... 42 3.7. CONCLUSIONS..................................................................................................................................... 44

4. EXPERIMENTAL ASSESSMENT OF REDUNDANT IMUS.............................................................45

4.1. INTRODUCTION.................................................................................................................................... 45 4.2. EXPERIMENTAL SYSTEM SETUP............................................................................................................. 45 4.3. IMPROVEMENT IN NAVIGATION PERFORMANCE..................................................................................... 46 4.4. NOISE REDUCTION AND ESTIMATION.................................................................................................... 48

4.4.1. Experimental Noise Reduction ...................................................................................................... 48 4.4.2. Assessment of the Noise Estimation Algorithm ............................................................................. 49

4.5. FAULT DETECTION AND ISOLATION......................................................................................................51 4.5.1. Assessment of the Noise Characteristics of in MEMS-IMUs......................................................... 51 4.5.2. Performance of the Parity Space Method with MEMS-IMUs ....................................................... 52

4.6. CONCLUSIONS..................................................................................................................................... 58

5. EMULATION OF REDUNDANT IMU SETUPS.................................................................................. 59

5.1. INTRODUCTION.................................................................................................................................... 59 5.2. INVERSE STRAPDOWN........................................................................................................................... 59 5.3. ERROR MODEL FOR INERTIAL SENSORS................................................................................................. 60 5.4. VARIANCE ANALYSIS............................................................................................................................. 62 5.5. EVALUATION OF THE NOISE CHARACTERISTICS OF INERTIAL SENSORS...................................................64

5.5.1. Quantization Noise........................................................................................................................ 64

3

5.5.2. White Noise ................................................................................................................................... 65 5.5.3. Bias Instability .............................................................................................................................. 66 5.5.4. Random Walk ................................................................................................................................ 67 5.5.5. Rate Ramp..................................................................................................................................... 68 5.5.6. Estimation of the Quality of the Allan Variance ........................................................................... 69

5.6. ESTIMATION OF THE PARAMETERS OF THE NOISE MODEL ..................................................................... 71 5.6.1. Estimation of the Scale Factor and of the Constant Part of the Bias............................................ 71 5.6.2. Estimation of the Stochastic Part of the Bias and of the White Noise Parameter ......................... 71

5.7. VALIDATION OF THE MODEL ................................................................................................................ 76 5.8. INFLUENCE OF THE SENSORS ORIENTATION IN REDUNDANT IMU .......................................................... 77 5.9. NAVIGATION PERFORMANCE IMPROVEMENT......................................................................................... 78 5.10. CONCLUSIONS..................................................................................................................................... 80

6. CONCLUSION AND PERSPECTIVES ................................................................................................. 81

6.1. SYNTHESIS........................................................................................................................................... 81 6.2. PERSPECTIVES..................................................................................................................................... 82

7. REFERENCES .......................................................................................................................................... 83

8. APPENDIX ................................................................................................................................................ 86

8.1. APPENDIX A: PROOF OF EQUATION (11)............................................................................................... 86 8.2. APPENDIX B: PROOF OF EQUATION (13)............................................................................................... 86 8.3. APPENDIX C: PROOF OF EQUATION (19) .............................................................................................. 87 8.4. APPENDIX D: PROOF OF EQUATION (23).............................................................................................. 88 8.5. APPENDIX E: PROOF OF EQUATION (42)............................................................................................... 89 8.6. APPENDIX F: PROOF OF EQUATION (60)............................................................................................... 90 8.7. APPENDIX G: PROOF OF EQUATION (72).............................................................................................. 92

4

List of Figures

Figure 1 : Two optimal configurations using five sensors. In the first case (left), the cone half-angle is 54. 74

[deg] (i.e. firstα ) and the sensors are separated by 72 [deg]. In the second case (right), the half-

angle is 65.91 [deg] (i.e. ondsecα ) and 90 [deg] separates the sensors. ........................................... 15 Figure 2 : The Platonic solids [13]................................................................................................................... 17 Figure 3 : Two minutes of static measurements with a MEMS-IMU (Xsens MT-i), after one minute vibrations

were added. The increase in the noise level after increased vibrations can be here observed ........ 19 Figure 4 : Theoretical noise reduction when using redundant sensors............................................................. 20 Figure 5 : Schematic representation of the averaging window ........................................................................ 21 Figure 6 : Graphical representation of the threshold-to-noise ratio – Redundancy - FAP .............................. 26

Figure 7 : Schematic representation of the threshold value T and 22 / nD σ probability density function

(pdf) ................................................................................................................................................. 27 Figure 8: Detector Operating Characteristics (DOCs) for various maximal errors (i.e. B) to 2

nσ ratio.......... 28

Figure 9 : Principle of mechanization based on a synthetic IMU .................................................................... 30 Figure 10 : Principle of extended IMU mechanization ...................................................................................... 33 Figure 11 : Principle of geometrically-constrained IMU mechanization ........................................................... 33 Figure 12 : Configuration of two sensors in planimetry .................................................................................... 35 Figure 13 : Configuration of three sensors in a three dimensional space........................................................... 36 Figure 14 : Influence of α and β on the information volume ........................................................................ 37 Figure 15 : Tetrad configuration proposed by A. Pejsa in [12].......................................................................... 39 Figure 16 : Influence of α and β on the information volume ........................................................................ 40 Figure 17 : Six sensors are placed in two different configurations. The symbol indicates the sensors that

will fail (superscript *). ................................................................................................................... 42 Figure 18 : Influence of λ and ϕ on the standard deviation of *2z ................................................................ 44 Figure 19 : (from left to right) (1) Skew-redundant IMUs placed in a tetrahedron. (2) Stéphane Guerrier

holding the scan2map system. (3) Installation of the base station by Yannick Stebler ................... 45 Figure 20 : Orientation errors after integration of a single sensor compared to................................................. 47 Figure 21 : Position, velocity, and acceleration errors after integration of a single sensor compared to extended

mechanization (four sensors)........................................................................................................... 48 Figure 22 : Comparison of the angular rate measurements of 4 MEMS-IMU to............................................... 49 Figure 23 : Estimation of the noise level for a selected portion of the experiment ............................................ 49 Figure 24 : Comparison between the noise estimates and the “real” noise (based on ln200’s data) calculated

during the experiment ...................................................................................................................... 50 Figure 25 : Orientation error after integration of four sensors (extended mechanization), with and without the

noise estimation algorithm............................................................................................................... 50 Figure 26 : Autocorrelation of errors of the Xsens MT-i’s gyros ....................................................................... 51 Figure 27 : Graphical assessment of the normality of Xsens MT-i errors (sensor 12, gyro ry).......................... 52 Figure 28 : Schematized algorithm for the detection and isolation of errors ..................................................... 53 Figure 29 : D values compare to T ................................................................................................................... 54 Figure 30 : Standford plot with TT ˆ= ............................................................................................................... 55 Figure 31 : D values compared to T value with identification of performance................................................ 55 Figure 32 : D values compared to

01.0T .............................................................................................................. 56 Figure 33 : Standford plot with

01.0TT = ............................................................................................................ 56

Figure 34 : D values compared to T values with identification of performance ............................................. 57 Figure 35 : Schematic representation of the Allan variance............................................................................... 63 Figure 36 : Allan variance of a white noise process........................................................................................... 65 Figure 37 : Example of two series of “bias instabilities” and white noise ......................................................... 66 Figure 38 : Allan variance of the series defined in Figure 37 ............................................................................ 67

5

Figure 39 : Allan variance of a random walk process........................................................................................ 68 Figure 40 : Schematic sample representation of Allan variance using analysis results ..................................... 69 Figure 41 : Typical Allan variances of MEMS gyroscope (computed with three hours static data); the spotted

lines represent the standard deviation of the Allan variance values. ............................................... 70 Figure 42 : Influence of the number of sample and the averaging time to the precision of the Allan variance 70 Figure 43 : Allan variance results for the gyroscopes (spotted lines indicate the standard deviation of the

measurements) ................................................................................................................................. 72 Figure 44 : Estimation of noise parameters for the gyroscopes (spotted lines indicate the standard deviation of

the measurements) ........................................................................................................................... 73 Figure 45 : Allan variance results for the accelerometers (spotted lines indicate the standard deviation of the

measurements) ................................................................................................................................. 73 Figure 46 : Estimation of noise parameters for the accelerometers (axis x) (spotted lines indicate the standard

deviation of the measurements) ....................................................................................................... 74 Figure 47 : Estimation of noise parameters for the accelerometers (axis y) (spotted lines indicate the standard

deviation of the measurements) ....................................................................................................... 74 Figure 48 : Estimation of noise parameters for the accelerometers (axis z) (spotted lines indicate the standard

deviation of the measurements) ....................................................................................................... 75 Figure 49 : Reference trajectory of an alpine skier (in Plain-Joux, Switzerland, experiment realized by Adrian

Wägli) .............................................................................................................................................. 76 Figure 50 : Comparison of the position errors with the simulated and real data................................................ 76 Figure 51 : Comparison of the velocity errors with the simulated and real data................................................ 77 Figure 52 : Comparison of the orientation errors with the simulated and real data ........................................... 77 Figure 53 : Position, velocity and orientations errors of two redundant (i.e. four sensors) systems placed

differently in space .......................................................................................................................... 78 Figure 54 : Influence of the redundancy of the position error and the maximal position error.......................... 78 Figure 55 : Influence of the redundancy of the velocity error and the maximal velocity error.......................... 79 Figure 56 : Influence of the redundancy of the orientation error and the maximal orientation error ................. 79

6

List of Tables

Table 1 : Average absolute accuracy of the tested GPS/MEMS-SRIMU system........................................... 46 Table 2 : Average and maximum errors of performance resulting from the use of the MEMS-SRIMU

compared to the performance of the individually integrated MEMS-IMU sensors......................... 47 Table 3 : Noise estimation for Xsens MT-i using Allan variance.................................................................... 75

7

Abstract

Nowadays, in many sports, differences in skills among competitors tend increasingly to

diminish and the margin between victory and defeat has become a matter of little details.

In this context, the Geodetic Engineering Laboratory of the Swiss Federal Institute of

Technology in Lausanne is developing a low-cost GPS/INS system for performance

analysis in sports. This master project is part of this larger project and aims to improve the

precision of this analysis using redundant IMUs.

In many sports, trajectory analysis is essential, and, as a result, competitive athletes have

turned to GPS techniques to help evaluate and improve their performance. This

technology, by recording athletes’ various positions, allows to analyze and compare

trajectories, velocities or accelerations. GPS satellite signals are, however, obstructed by

various obstacles in athletes’ environments, therefore making the resolution of phase

ambiguities difficult or even impossible. To overcome these difficulties, Inertial

Measurement Units (IMUs) are integrated with GPS, which also enables a precise

calculation of accelerations and orientations. For ergonomic and economical

considerations, the systems considered here are composed of MEMS-IMUS together with

low-cost L1 GPS receivers.

The research for this project has demonstrated the following points. Firstly, redundant

inertial sensors enhance navigation performances. With four sensors, for example, errors

are diminished by 30-50%. Secondly, optimal sensor geometry can be formalized and

proves that, in most cases, the performance of the system is independent of this geometry.

Thirdly, the measurements’ redundancy reduces the noise affecting observations, as well as

estimates variations in the noise level. In addition, the MEMS IMU errors can be detected,

isolated and corrected successfully. Finally, measurements of MEMS sensors can be

simulated with an appropriate error model. Consequently, navigation performances of

redundant systems can be theoretically estimated though simulation.

8

Résumé

Actuellement, dans le monde du sport, les écarts entre athlètes tendent petit à petit à

s’amoindrir et bien souvent la différence entre une victoire et une défaite trouve son

explication dans quelques petits détails. Dans ce contexte, le laboratoire de Topométrie de

l’Ecole Polytechnique Fédérale de Lausanne (EPFL) est actuellement en train de

développer un système GPS/INS dédié aux analyses des performances sportives. Ce travail

de master s’intègre dans le cadre de ce grand projet et s’intéresse plus particulièrement à

l’utilisation de capteurs inertiels redondants afin d’améliorer la précision de l’analyse de

ces performances.

Dans tous les sports où l’étude des trajectoires est essentielle, les athlètes s’intéressent au

GPS afin d’évaluer leurs performances. En effet, celui-ci enregistre des postions qui

permettent d’analyser puis de comparer des trajectoires, des vitesses ou des accélérations.

Mais, fréquemment l’environnement des athlètes ne permet pas une utilisation performante

de ce système. Ainsi, il est nécessaire de le coupler à des mesures inertielles qui, de plus,

permettent un calcul précis des accélérations et des orientations, déterminantes dans

certains sports. Pour des raisons économiques et ergonomiques, les systèmes considérés

dans ce travail sont basés sur des capteurs inertiels de type MEMS, ainsi que des

récepteurs GPS monofréquence.

Ce travail permet de mettre en évidence les éléments suivants. Tout d’abord, les capteurs

inertiels redondants améliorent les performances de navigation. Avec quatre capteurs, par

exemple, l’erreur est diminuée de 30-50%. Les considérations liées à la géométrie des

systèmes redondants peuvent être formalisées et montrent que, dans la majorité des cas, la

performance du système est indépendante de cette géométrie. Ensuite, la redondance de

mesures permet de réduire le bruit qui l’entache, celle-ci permet, de plus, une estimation

des niveaux de bruit. De surcroit, les erreurs affectant les mesures des capteurs MEMS

peuvent être détectées, isolées et corrigées améliorant, ce faisant, les performances de

navigation. Enfin, les mesures de capteurs MEMS peuvent être simulées grâce à un modèle

d’erreur adapté. Ainsi, les performances de navigation d’un système redondant peuvent

être théoriquement estimées par simulation.

9

Remerciements

Je tiens à remercier ici toutes les personnes qui m’ont aidé à réaliser ce projet mais aussi

toutes celles qui m’ont permis d’effectuer ce parcours.

Tout d’abord, mon professeur, Dr. Jan Skaloud, qui a su me passionner, m’encourager et

me donner l’envie de poursuivre dans cette voie. Je tiens également à le remercier pour sa

grande disponibilité, son aide dans le cadre de ce projet et ses conseils judicieux mais aussi

pour la confiance qu’il m’a accordée en me laissant beaucoup de liberté.

Je tiens à remercier tout particulièrement Adrian Wägli, sans l’aide duquel ce projet

n’aurait pu être réalisé. Nos échanges très intéressants, le plaisir de travailler avec

quelqu’un d’aussi brillant qui, de plus, allie la détermination à l’humour furent une chance

pour moi. Sans oublier tous les bons moments passés ensemble.

Ensuite, je souhaiterais remercier le Prof. Bertrand Merminod, pour son enseignement de

qualité qui a éveillé mon intérêt pour ces disciplines et qui a toujours pris le temps de

répondre à mes questions.

Le Prof. Ishmael Colomina pour son aide, l’intérêt de ses travaux et son hospitalité. Ainsi,

qu’Eulàlila Parés, Pere Molina Mazón, Marino Wis et les collaborateurs de l’Institut de

Geomàtica qui m’ont si bien accueilli à Barcelone et avec qui j’ai eu beaucoup de plaisir à

travailler.

Un grand merci à Yannick Stebler, mon camarde d’étude, pour son esprit vif, les échanges

constructifs que nous avons eu et notre chaleureuse complicité.

J’aimerais aussi remercier les collaborateurs du laboratoire de Topométrie pour la bonne

ambiance et l’esprit d’équipe qui y règnent.

Le Prof. Hannelore Lee-Jahnke pour ses amicaux et judicieux conseils à la rédaction de ce

travail en anglais, langue dont je suis loin de maîtriser toutes les subtilités.

Enfin, Justine et Cédric qui ont pris le temps de relire ce travail ainsi que pour tous les

bons moments que nous avons passés ensemble et leur précieuse amitié.

Finalement, je tiens à remercier mes parents qui m’ont toujours soutenu et aidé. Sans

lesquels je n’aurais pu en arriver là !

10

Abbreviations

EKF Extended Kalman Filter

FDI Fault Detection and Isolation

GPS Global Positioning System

IMU Inertial Navigation System

INS Inertial Navigation System

MEMS Micro-Electro-Mechanical System

PDF Probability Density Function

PSD Power Spectral Density

RMS Root Mean Square

SRIMU Skew Redundant Inertial Navigation System

SSE Sum of Squared Errors

11

1. Introduction

1.1. Context of the Study

This master project was realized in the Geodetic Engineering Laboratory of the Swiss

Federal Institute of Technology Lausanne (EPFL). It is part of a larger investigation which

aims at developing a low-cost GPS (Global Positioning System) / INS (Inertial Navigation

System) system for performance analysis in sports [1-4].

Nowadays, in many sports the differences of skills between competitors tend increasingly

to diminish. Consequently, training for competition, in which every detail counts, has

become a key factor of success. Traditionally, athletes’ performances have been analyzed

using chronometry or video recordings. However, these techniques are limited by

meteorological conditions and by the difficulty of replicating postures and movements

across trials. Moreover, these only provide few quantitative variables that could be used to

analyze objectively athletes’ performances. New methods are, on the contrary, being

sought that offer precise measurements of positions, velocities, accelerations and forces

[4]. Indeed, satellite-based positioning has already been successfully applied in many

sports such as skiing [2, 4], car racing [5] or rowing [6]. Taking this into consideration,

athletes’ environments are often partially composed of areas that may block or attenuate

satellite signals and therefore make the resolution of phase ambiguities difficult or even

impossible. To overcome these difficulties Inertial Measurement Units (IMUs) are

integrated with GPS, which also enable an accurate determination of accelerations and

orientations. Indeed, in numerous applications the orientation is required. For example, in

motorcycling the correct exploitation of torque and force sensors necessitates knowledge

of the sensors’ orientation. Additionally, the attitude determination also provides essential

data to study the vibratory behavior of the pneumatics [3].

Traditionally, GPS/INS equipment consisting of dual-frequency GPS receivers and

tactical-grade INS, provides high accuracies even for large dynamics (cm for position,

cm/s for velocity and 1/100° for orientation). However, such equipment is bulky (a few kg)

and expensive ( > 40’000€) and is consequently not suitable for most sport applications

[3]. Recently, for ergonomic and economical considerations, smaller and lighter

equipment, composed of low-priced Micro-Electro-Mechanical System (MEMS) IMUs

together with low-cost L1 GPS receivers, were successfully applied in alpine skiing and

motorcycling [3, 4, 7, 8].

The outline of this work is to improve navigation performance (and more specifically

orientation) by using redundant MEMS-IMUs with GPS. Indeed, position and velocity can

12

be improved by using dual-frequency GPS receivers, instead of single-frequency receivers.

This, however, cannot be done for orientation. The assessment of the accuracy of this

technology was measured in the field of sports, the reach of this project, however, goes far

behind the sports world. In fact, in many fields, redundant MEMS-IMUs are already used

(e.g. robotics, virtual reality). Above all, this technology could be implemented in a wide

range of domains (e.g. low-cost reliable navigation systems, pedestrian navigation).

Unfortunately, MEMS-IMU are prone to large systematic errors (e.g. biases, scale factors,

drifts), which limits their applicability in integrated navigation systems [1]. Typically, the

performance of these IMU reaches a position accuracy of 0.5 m, a velocity accuracy of 0.2

m/s, as well as an orientation accuracy of 1 deg for the pitch and roll and 2 deg for the

heading. The sensors used are highly miniaturized and thus offer the possibility of using

numerous IMUs to enhance the determination of attitude. The redundancy of

measurements can improve the performances of the GPS/INS integration at several levels.

Firstly, direct noise estimation can be achieved directly from the data in order to improve

the stochastic model of the Extended Kalman Filter (EKF). Secondly, the noise level can

be reduced and defective sensors, spurious signals and sensor malfunctioning can be

detected and isolated. Finally, sensor error calibration becomes conceivable even during

uniform motion or static initialization. Due to the improved orientation accuracy,

redundant IMUs may bridge the gap in the GPS data more effectively [1].

Redundancy in inertial navigation has already been investigated with higher-order IMUs in

photogrammetry and remote sensing [9]. Several authors have presented results with

MEMS-IMU based on simulations or emulations and have found an accuracy improvement

of 33% with MEMS-IMUs placed on a tetrahedron [10]. Emulations presented by [11]

resulted in performance improvements of 20-34%. The first experimental results with

MEMS-MUs were given in [1] and have shown an improvement of navigation

performance, when using four MEMS-IMUs, of 30-50%.

1.2. Objectives

This research project focuses on the potential benefits of using redundant MEMS-IMUs to

improve GPS/INS integration. The objectives are:

• Determine optimal sensor architecture (i.e. geometry)

• Detect and isolate defective sensors and spurious signals

• Reduce noise and direct noise estimation of sensors

• Calibrate sensor even during uniform motion and initialization

13

• Assess the experimental navigation improvements stemming from the use of

redundant MEMS-IMUs.

• Develop simulation tools in order to assess theoretical improvements of redundant

configurations

1.3. Structure of the Rapport

This rapport is structured in four parts.

The first part presents the theoretical basis of IMU redundancy. It details how optimal

sensor architectures can be determined, and describes how noise is reduced and can be

estimated when redundant inertial sensors are employed. In addition to that, it presents one

of the most commonly used algorithm of Fault Detection and Isolation (FDI) and different

mechanization’s methods which can be used with redundant IMUs.

The second part studies sensor geometries and the influence of this geometry on the system

performance.

The third part assesses experimentally the advantages of using multiple IMUs. It shows the

noise reduction as well as its estimation and the subsequent improvements brought to the

navigation solution (i.e. position, velocity and orientation). Finally, the use of redundant

IMUs allows to detect and isolate erroneous measurements. In this study, the performance

of the parity space method with MEMS-IMU is presented.

The last part presents the development of a simulation tool to emulate MEMS-IMU

measurements. It details the used error model and explains how its parameters are

estimated. As a conclusion, the theoretical improvements of various redundant

configurations are evaluated.

14

2. Theoretical Basis of IMU Redundancy

2.1. Introduction

This chapter aims to present the theoretical basis of IMU redundancy. It covers

geometrical considerations about the spatial configurations of inertial sensors, as well as

the theoretical basis of noise reduction achieved with sensor redundancy. This chapter also

presents an algorithm that estimates the evolution of noise during the processing. In

addition, it contains a review of the most common FDI algorithm, the parity space

approach. Finally, different mechanization methods which can be applied with redundant

IMUs are presented.

2.2. Geometrical Configurations of Inertial Sensors

Redundant IMUs have been used since the early days of the inertial technologies in safety

critical operations such as in the control of military or space aircrafts. These aircrafts are

designed to be dynamically unstable, in order to increase manoeuvrability, while inertial

accelerations and rotation rates are used to observe the vehicle’s stabilizing parameters

[10]. Clearly, such systems require sensor redundancy. The redundant information was

used to create fault-tolerant systems which were able to detect and isolate defective sensors

(i.e. FDI). If two sensors are placed collinear to each other, it is possible to detect a fault

that occurs in either one of the sensors. To isolate the erroneous device at least three

sensors are required. Therefore, traditionally, nine sensors were used in a full three-

dimensional system (three per axis).

In 1974, [12] proposed a first theory to optimally position any number of sensors which

showed that less than nine sensors (four in theory) were required to isolate faults in a three-

dimensional space. This theory essentially considers two situations: firstly, when sensors

are equally spaced in a cone of half-angle α and, secondly, when one sensor is placed

along the central cone axis while the remaining sensors are positioned equally around a

cone of half-angle α . The optimal half-angle α corresponds to the configuration in which

the variance is minimized. The results for α obtained with the two configurations are

shown in Figure 1. Numerically, it yields to:

15

( )

−⋅−=

=

33

3arccos

arccos

sec

33

n

nond

first

α

α (1)

where n is the number of sensor.

xy

z

90 deg.

α

xy

z

72 deg.

α

Figure 1 : Two optimal configurations using five sensors. In the first case (left), the cone half-angle

is 54. 74 [deg] (i.e. firstα ) and the sensors are separated by 72 [deg]. In the second case (right),

the half-angle is 65.91 [deg] (i.e. ondsecα ) and 90 [deg] separates the sensors

2.2.1. Information Filters

Years later, [10] proposed an other method using information filters and showed that it

could be used to maximize the amount of information in redundant sensors configurations.

This approach provides a better conceptual understanding of how the information is

distributed in space.

In the state space, the observation at time k is given by:

( ) ( ) ( ) ( )kkkk vxHz +⋅= (2)

where ( )kx is the current state, ( )kH is the observation model and ( )kv is the

observation noise with covariance ( )kR .

The information contribution to the states of an observation constitutes the information

observation vector ( )ki defined as:

16

( ) ( ) ( ) ( )kkkk T zRHi ⋅⋅= −1 (3)

The amount of certainty associated with this observation is given by the information

matrix,

( ) ( ) ( ) ( )kkkk T HRHI ⋅⋅= −1

(4)

Note that the matrix ( )kI is independent from the observations, since ( )kz is absent from

the equation. Thus, it provides a measure of the certainty of these observations purely

based on the geometry of the system. Moreover, it is often assumed that all sensors are

uncorrelated and of equal variance. Under this assumption, equation (4) can be rewritten

as:

( ) ( ) ( )kkk T HHI ⋅⋅= 20

(5)

Since the information matrix is positive and semi-definite, its corresponding eigenvectors

are orthogonal to each other. These eigenvectors can, additionally, be seen as the axes of

an ellipsoid in the information space. If the eigenvalues of those eigenvectors are different,

then the larger axis of the ellipsoid represents the direction in which information is

maximal. Thus, by maximizing the volume of the ellipsoid, the volume of information is

also maximized. Since the determinant of the information matrix is directly related to the

volume of the ellipsoid, the goal is to maximize the expression presented in equation (6).

( ) ( ) ( )( )( )kkJ T HH ⋅= detmaxmax (6)

The maximization of J shows that the optimal configuration adopts the shape of a sphere

in the information space. Hence, in this optimal geometry, the amount of information place

for each axis is equal.

This approach was applied to 4, 6, 8, 12 and 20 sensor combinations. The corresponding

optimal configurations are regular polyhedrons (tetrahedron, cube, octahedron,

dodecahedron and icosahedron respectively) which verifies the results of [12]. These solids

are perfectly symmetrical and are known as the Platonic Solids (see Figure 2).

17

Figure 2 : The Platonic solids [13]

Flowing from these results, when the configuration is optimal and regardless of the number

of sensors, the eigenvectors of the information matrix ( )kI are equal to the number of

sensors divided by the number of dimension. Thus,

=D

niλ (7)

where n is the number of sensors, D the number of dimensions and 3 2, ,1=i . This

equation suggests that, in the information space, each sensor contributes to a third of its

information to each eigenvectors. Moreover, equation (7) enables to quantify the

maximum volume of information of n sensors placed in an optimal configuration as the

product of the eigenvalues, that is,

3

max

=D

nJ (8)

In short, this approach, based on information filters, allows to analyze optimal

configuration and provides a geometric visual understanding of the distribution of

information across axis. Moreover, it can also be used to create geometries with specific

uneven distributions of information (e.g. more information in planimetry than in altimetry).

2.2.2. Partial Redundancy

Another interesting approach to determine optimal sensor configurations is based on the

concept of reliability. This method is typically applied in geodetic networks and has – to

the author’s knowledge – never been applied in a context similar to that of this study.

18

Indeed, in geodetic networks with redundant observations, the measurements control each

other. In order to assess the “amount” of controllability of each measurement, a iz value

can be computed and associated to every observation. This value, called partial

redundancy, corresponds to the participation of the observation i to the global redundancy

of the system. It varies from 0 to 1, an observation having an associated 0=iz isn’t

controlled at all. To the contrary, a 1=iz implies a “totally” controlled observation. In

geodetic engineering, a network is typically considered good (i.e. technically and

economically) if all measurements have an associated ( )0.25,0.60 ∈iz [14].

To compute the partial redundancy, we first calculate the Z matrix is defined as:

( ) 1−⋅⋅⋅−= RHPHRZ T (9)

where R is the covariance matrix of the observations, H is the design matrix. P is the

covariance matrix of the state vector, defined as:

( ) 11 −− ⋅⋅= HRHP T

(10)

The partial redundancies iz are the diagonal elements of the Z matrix. Note that

equations (9) and (10) show that iz depends solely on the geometry of the system (i.e.

design matrix H ) and on the accuracy of the measurements (i.e. covariance matrix R ).

Moreover, it has been shown that:

∑=

=−=n

iizunr

1

(11)

Where r is the system redundancy, n is the number of measurements and u the number

of parameters. Equation (11) is demonstrated in the Appendix A.

Thus, the geometry of the system will influence the iz but not their summation.

Consequently, an optimal sensor configuration would minimize the differences between

these measurements to approach the optimal case, where observations are equally

controlled. Mathematically, this is equivalent to minimizing of the standard deviation of

the iz , that is:

Best configuration(s): [ ][ ]2 min ii zEzE − (12)

19

2.3. Noise Reduction and Estimation

Combinations of redundant inertial sensors not only decrease noise measurements, but also

offer the possibility to estimate, with the help of adaptive filters, the level of noise during

the processing. Indeed, the noise level can evolve during the processing in response to

particular situations (e.g. increased vibrations). Figure 3 shows an example of such

behaviour.

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Time [s]

Ang

ular

rot

atio

n ra

te

[rad

/s]

Figure 3 : Two minutes of static measurements with a MEMS-IMU (Xsens MT-i), after one minute vibrations were added. The increase in the noise level after increased vibrations

can be here observed

When redundant IMUs are used, it is possible to quantify by how much redundancy

reduces the noise. Theoretically, from n independent measures 21 , ... , xx (with their

respective variances 221 , ... , nσσ ), we can compute their best estimate x . Assuming

homogenous measurements (i.e. constant iσ ), its variance 2xσ can be derived as [15, 16]:

nw x

n

iiix

σσσ == ∑=1

22ˆ

(13)

where the iw are weighting factors and iix ∀= σσ . The derivation of equation (13) is

given in Appendix B. This theoretical noise reduction corresponding to various redundant

IMUs configurations is represented in Figure 4

20

1 2 3 4 5 6 7 8 9 10 11 120

10

20

30

40

50

60

70

80

90

100

# of sensors

Noi

se le

vel [

%]

Figure 4 : Theoretical noise reduction when using redundant sensors

Another interesting feature of multiple IMUs is the possibility of using a noise estimation

algorithm during the processing. This information can be used to improve the performance

of the GPS/INS integration by providing a more realistic stochastic model. Moreover, it

can improve the FDI performances which depend on the noise model.

The algorithm we propose to use in this case is laid out below.

Firstly, the norms, which is not sensible to the orientation, of angular rotation rates and

specific forces are computed for every time k and for every sensor i :

( ) ( ) ( ) ( )

( ) ( ) ( ) ( )kfkfkfk

krkrkrk

zyx

zyx

iiii

iiii

222

222

++=

++=

f

r (14)

where ni ,,1K= and n is the number of sensors triads.

Squared differences are then computed to form the d matrices:

21

( )

( )( )

( )( )

−−

−−

=

=

=

21

1

2

21

1

1

,

n

k

k

n

k

k

knn

jj

n

n

jj

r

r

r

r

d r M and ( )

( )( )

( )( )

−−

−−

=

=

=

21

1

2

21

1

1

,

n

k

k

n

k

k

knn

jj

n

n

jj

f

f

f

f

d f M (15)

An averaging time T (typically a few seconds, see Figure 23) is chosen (see Figure 5).

Assuming all sensors’ errors to be uncorrelated and identical, the noise is directly

estimated:

( ) ( )

( ) ( ) dtkin

Tk

dtkin

Tk

Tk

Tk

n

if

Tk

Tk

n

ir

=

=

∫ ∑

∫ ∑

+

−=

+

−=

1

1

,1

,

,1

,

f

r

d

d

σ

σ

(16)

T TT T

( )k time

fr dd /

Figure 5: Schematic representation of the averaging window

22

2.4. Fault Detection and Isolation

2.4.1. Introduction

Fault Detection and Isolation (FDI) algorithms for inertial navigation have been thoroughly

investigated in the past. The most commonly used approach is the parity space method [10,

17-20], but other approaches such as artificial neural networks have also been employed

[21]. The complexity of implementation of an efficient FDI system is increased using

MEMS-IMUs. Indeed, their poor performance (i.e. noise density variations and larger

proportion of systematic errors (colored noise, bias, scale factor) compared to the higher

grade-IMU) creates high risks of false alarm, as well as an increased exposure to

misdetection of faulty measurements [10]. We investigate here the parity space method

with skewed redundant MEMS-IMUs. In the context of this project, focused on non-safety

critical sports applications, the main objective is to increase navigation performances and

to prevent gross errors.

2.4.2. The Parity Space Method

2.4.2.1. Measurement Model

In general, we consider l measurements in an n dimensional state space which is over

determined (i.e., 1+≥ nl ). The system can be defined by equation [20]:

bnxHz ++⋅= (17)

where z is the ( )1×l measurement vector compensated with all prior information; H is

the ( )nl × design matrix which converts the state space into the measurement space; x is

the ( )1×n state vector; n is a ( )1×l vector of Gaussian white noise having a variance of

2nσ and b is a ( )1×l vector of uncompensated measurement biases (i.e. faults).

Consider the case of m IMUs, equation (17) can be rewritten as:

bn

m

f

w

A

A

A

l

l

l

++

=

oo

oo

oo

b

b

b

m

f

w

bm

bf

bw

i

i

i

(18)

23

where ibwl , ib

fl and ibml represent the observation vectors (respectively for the gyroscopes,

accelerometers and magnetometers), bw , bf and bm represent IMU measurements in the

pre-define body frame of reference. Furthermore, wA , fA and mA are the matrices that

transforms from the reference body frame (superscript b ) into the individual sensor frames

of the IMUs (superscript ib ) (these matrices are defined in equation (38)).

The fault model in this case is a single measurement source failure which results in a step

bias shift. A failure of the thi measurement is modeled by ibb = where ib is a ( )1×l

vector with the magnitude of the error B at the thi position, zero elsewhere. If no faults

occurs then 0b = [20].

Additionally, equation (17) shows that the underlying assumption behind the parity space

method is that non erroneous measurements are only subject to Gaussian white noise

errors. This assumption is acceptable in the case of a calibrated high performance IMU,

but may not realistic with MEMS-IMU which are subject to colored noise.

2.4.2.2. Parity Space Model

The parity space method’s theoretical basis presented here are largely based on [20].

The least square best estimate of the state vector of the equation (17) is given by:

zHx * ⋅=ˆ where ( ) T1T* HHHH ⋅⋅= − (19)

The derivation of this equation is given in Appendix C.

Thus, *H converts measurements space into state space. For any ( )nl × H design matrix

with a rank n , a ( )( )nnl ×− P matrix can be found such that the rank of P is ( )nl − ,

nlT IPP −=⋅ and 0HP =⋅ . So, the matrix P spans the null space of H (i.e. the parity

space).

Moreover, the ( )ll × matrix

=

P

HA

*

can be formed and has a rank l . This matrix

represents a linear transformation of the measurement space into two subspaces, namely

the state space and the parity space.

The parity vector is calculated in the following way:

24

( ) ( )bnPbnPxHPzPp +⋅=+⋅+⋅⋅=⋅= (20)

Thus, this vector is independent from x , its elements are normally distributed and have an

expected value of: [ ] ( )[ ] [ ] [ ] bPbPnPbnPp ⋅=⋅+⋅=+⋅= EEEE . What’s more, the

covariance of the parity vector is defined as: [ ] nlnσCOV −⋅= Ip 2 . Consequently, the

elements of p have an equal variance (i.e. the noise variance) and are not correlated. If no

fault occurs, then 0b = and [ ] 0p =E .

The matrix [ ]TPHA 1 =− (see Appendix D) also has a rank l and represents the inverse

transformation caused by A , therefore it transforms the state and parity spaces into a

measurement space. The conversion of the parity vector into the measurement space is

obtained by:

[ ] zSzPPp

0PHf ⋅=⋅⋅=

⋅= × TnT 1 (21)

where f is the fault vector having the following expected value:

[ ] [ ] [ ] [ ] [ ] bSbSnSxHPPzSf ⋅=⋅+⋅+⋅⋅⋅=⋅= EEEEE T (22)

The matrix PPS T ⋅= can be directly computed from H using the equation:

*l HHIS ⋅−= (23)

The derivation of this equation is given in Appendix D.

2.4.2.3. Fault Detection Algorithm

The first step in any FDI system is to detect erroneous measurements. This step falls within

the general framework of composite hypothesis tests [17]. The fault detection can be

viewed as a choice between two hypotheses, in the absence or in the presence of a faulty

measurement. The hypothesis test is based on a comparison of a decision variable D to a

threshold variable T . The test can be defined as:

25

⇒>⇒<

detected isfault a :H

fault no :H

1

0

TD

TD

where 0H is the null hypothesis and 1H the alternative hypothesis.

The performance of the test is defined by the probabilities of false alarm FAP and of

misdetection MDP , which are expressed by:

( ) ( )( ) ( )iMDMD

FAFA

TDPPTDPP

TDPPTDPP

bb

0b

=<=⇔<=

=>=⇔>=

1

0

H

H (24)

The decision variable D is considered to be the sum of the square magnitude of the fault

vector (i.e. SSE) [17, 20, 22]. Thus, D is defined as:

ff ⋅= TD (25)

The definition of the threshold variable T is, however, more complicated. In the fault-free

case 0b = , [ ] 0p =E and 2/ nD σ has a chi-square distribution (since it is a summation of

normal distributions [23]) with nlr −= degrees of freedom [17, 20, 22, 24]. This implies

that:

= r

TQP

nFA 2σ

(26)

where ( ) ( )rPrQ 22 1 χχ −= and ( )rP 2χ is a chi-square probability function defined as:

( ) ( )[ ] dtetrPtrr

r 2

2

22

2

0

1

22 2 −−

∫−

⋅Γ=χ

χ (27)

Consequently, the probability of false alarm FAP depends on the number of redundant

measurements r and is independent of the design matrix H .

The required threshold can now be determined for any given variables FAP , r , 2nσ using:

( ) ( )nlPQnlPT FAnnFA −⋅⋅=− −122,, σσ (28)

26

where ( )rQ 21 χ− is the inverse function of ( )rQ 2χ .

Thus, there is a relation between the ratio threshold-to-noise, the variance ratio 2/ nT σ , the

probability of false alarm FAP and the redundancy of the system r . Figure 3 shows this

relationship.

00.05

0.10.15

0.2

0

5

10

15

0

5

10

15

20

25

30

35

40

PFA

Redundancy

Thr

esho

ld-t

o-N

oise

Rat

io (

T/

σσ σσn2 )

2.5

5

5

7.5

7.5

1 0

10

12.

5

12.5

12.5

15

15

15

17.

517

.5

17.5

20

20

20

22.

52 2

. 52

52 5

PFA

Red

unda

ncy

0.05 0.1 0.15 0.2

2

4

6

8

10

12

0

5

10

15

20

25

Thr

esho

ld-t

o-N

oise

Rat

io (

T/σσ σσ

n2 )

Figure 6: Graphical representation of the threshold-to-noise ratio – Redundancy - FAP

Further, it is assumed that faults are equally likely to occur in any measurement, which

implies that:

( )∑=

=<⋅=l

iiMD TDP

lP

1

1bb (29)

But if ibb = then [ ] iE bPp ⋅= and 2/ nD σ has non-central chi-square distribution with r

degrees of freedom and a non-centrality parameter defined by [17, 20, 22, 24], then:

iin

i SB ⋅

= 2

2

σθ (30)

27

where iiS represents the thi element of the diagonal of the matrix S (defined in equation

(23)). Hence:

∑=

−=

l

iii

nnMD S

Bnl

TP

lP

12

2

2,

1

σσ (31)

where ( )θχ ,2 rP is a non-central chi-square probability function defined by:

( ) ( ) ( )jrPj

erPj

2!

, 222 2 += − χθχθθ

(32)

Figure 7 shows schematically the relationship between the threshold value T and the

central or non-central 2χ probability distribution function of 22 / nD σ .

PFA

PMD

D2/σσσσ2n (faulty case)

Non-central χχχχ2 pdf

D2/σσσσ2n (fault free case)

central χχχχ2 pdf

Threshold

Figure 7 : Schematic representation of the threshold value T

and 22 / nD σ probability density function (pdf)

The probabilities of false alarm FAP and of misdetection MDP having been explained, the

Detector Operating Characteristics (DOCs) can now be defined. The DOCs are the

graphical representation of MDP vs FAP [20]. In some application a certain MDP is required,

however, the value of the threshold is based solely on FAP . Hence, the DOCs allow to

28

determine from any FAP the corresponding MDP (Figure 8). Equation (33) describes this

relationship.

( ) ( ) ( )∑∑=

=

−−⋅=

−=

l

iii

nFA

l

iii

nn

FAFAMD S

BnlnlPQP

lS

Bnl

PTP

lPP

12

21

12

2

2,

1,

1

σσσ (33)

Figure 8 shows examples of DOCs:

10-3

10-2

10-1

100

10-3

10-2

10-1

100

PFA

PM

D

B/σn = 5

B/σn = 7.5

B/σn = 10

B/σn = 12.5

B/σn = 15

B/σn = 17.5

B/σn = 20

Figure 8: Detector Operating Characteristics (DOCs) for various maximal errors (i.e. B) to 2

nσ ratio

2.4.2.4. Fault Isolation Algorithm

The fault identification process is based on the maximum likelihood estimation (MLE)

approach [17]. The isolated erroneous measurement is the value having the highest iii /Sf 2

ratio [20]. Moreover, the probability of misidentification, under the assumption of equally

likely measurement faults is:

<= ibMAX

jj

j

jii

iMI S

f

S

fP

lP

221 (34)

29

2.5. Integration of Redundant IMUs with GPS

This section aims to present the mechanization approach applied for the integration of GPS

with a single IMU. Then, it exposes the three mechanizations approaches described by [9].

2.5.1. Single IMU Mechanization

An extended Kalman filter (EKF) has been implemented in the local level frame

(superscript n ) which makes the interpretation of the state variables straightforward. The

following strapdown equations needs to be solved [1, 15]:

( )( )

+⋅+×+−⋅=

nin

nib

nb

nnnie

nin

bnb

n

nb

n

n

ωωR

gvωωfR

v

R

v

r

&

&

&

(35)

Given the observations bibω and bf , the knowledge of ng , and the initial conditions.

For the inertial measurements, a simplified model was considered. Indeed, assuming that

the misalignments, drifts and constant offsets could not be decorrelated efficiently given

the characteristics of the MEMS-IMU sensors and limited integration periods, only a bias

term is considered [1, 8]. Their associated errors are modeled as first order Gauss-Markov

processes:

bb ll

bb ll wb ++=ˆ

bbbbbb llllllwbb ⋅⋅⋅+⋅−= βσβ 22&

(36)

where bl is the estimated inertial observation (specific force bf or rotation rate bibw ), bl

the inertial measurement, blb the bias of the inertial measurement, blw the measurement

noise, 2bl

σ the covariance at zero time lag and blβ the inverse of the correlation time [1,

25].

2.5.2. Synthetic Mechanization

In the synthetic mechanization approach, redundant inertial data are projected to an

arbitrary non-redundant IMU (i.e. the synthetic IMU), before being introduced in the

30

GPS/INS algorithm based on the standard single IMU mechanization (Figure 9) [1, 9]. In

this section, the mathematical construction of a synthetic IMU composed of n triad of

accelerometers, gyroscopes and magnetometer will be described.

IMU 1IMU 1

IMU 2IMU 2

IMU nIMU n

GPSGPS

Synth. IMUSynth. IMU NAV solNAV solNAV procNAV proc

Figure 9 : Principle of mechanization based on a synthetic IMU

Assume the vectors ibwl , ib

fl and ibml represent the observation vectors for the gyroscopes,

the accelerometers and the magnetometers respectively. For example, lwl can be described

as:

[ ]Tzyxzyxlw nnn

rrrrrr K111

=l

Then, assuming perfect measurements it follows that:

=

b

b

b

m

f

w

bm, true

bf, true

bw, true

oo

oo

oo

i

i

i

m

f

w

A

A

A

l

l

l

(37)

where bw , bf and bm represent the synthetic IMU “measurements.” wA , fA and

mA are ( )33 ×⋅n matrices that transform the reference body frame (superscript b ) into

the body frames of the sensor triads (superscript ib ). These transformation matrices can be

view as the product of two rotation matrices that can be expressed as:

31

),(λ

),(λ

),(λ

nnnj

j

j

j,j

jj

jj

b

b

b

b

b

b

j

=

=

ϕ

ϕϕ

R

R

R

R

R

R

AMM

22

11

2

1

(38)

where mfwj ,,= and where:

( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( )

( ) ( )

−⋅−⋅−⋅−−⋅−

=λλ

λλ

λλ

)(λ

sin0cos

sincoscossinsin

coscossincossin

, ϕϕϕϕϕϕ

ϕR

(39)

Moreover, equation (37) can be rewritten as an error equation by introducing residual

vectors i

w

blv , i

f

blv and i

m

blv , which yields to:

oo

oo

oo

b

b

b

m

f

w

bl

bl

bl

bm

bf

bw

i

m

i

f

i

w

i

i

i

=

+

m

f

w

A

A

A

v

v

v

l

l

l

(40)

Hence, equation (40) transforms the redundant observation vectors into the observation of

the “synthetic” IMU. Now, to determine the synthetic measurements equation (40) has to

be solved for bw , bf and bm . Considering the redundant nature of the problem, the well-

known least-squares estimation seems appropriate (see Appendix C). This yields to:

=

bm

bf

bw

m

f

w

b

b

b

oo

oo

oo

l

l

l

Π

Π

Π

m

f

w ( )( )( )

RAARAΠ

RAARAΠ

RAARAΠ

⋅⋅⋅⋅=

⋅⋅⋅⋅=

⋅⋅⋅⋅=

−−−

−−−

−−−

111

111

111

where

bm

bm

bm

bm

bf

bf

bf

bf

bw

bw

bw

bw

ll

Tmmll

Tmm

ll

Tffll

Tff

ll

Twwll

Tww

(41)

where bw

bwll

R , bf

bf ll

R and bm

bmll

R are the covariance matrices of the raw measurements vectors

ibwl , ib

fl and ibml respectively. Under the assumption IR ⋅= 2

0σbw

bwill

, equations (41) can be

simplified as follows:

32

=

=

=

⋅⋅=

⋅⋅=

⋅⋅=

n

iz

y

xTb

bb

n

iz

y

xTb

bb

n

iz

y

xTb

bb

i

i

i

im

i

i

i

if

i

i

i

iw

m

m

m

n

f

f

f

n

r

r

r

n

1

1

1

1

1

1

Rm

Rf

Rw

(42)

These equations are derived in Appendix E.

This synthetic approach does not require modifications of the GPS/INS software, because

the synthetic IMU can be considered as a single IMU. Futher, while fusing the IMU data,

defective sensors can be detected and covariance terms can be estimated. The estimated

“compound” biases, however, cannot be back projected to the individual sensor space and

the noise may be inflated by these unknown errors [1, 9].

2.5.3. Extended Mechanization

The extended mechanization estimates individual sensors errors. Thus, equations (35) and

(36) have to be adapted in order to take into account the sensor redundancy [1, 9]:

( )( )

−⋅⋅+×+−⋅⋅=

ninw

nb

nnnie

nin

nbf

n

nb

n

n

l

l

ωΠR

gvωωRΠ

v

R

v

r

w

f

&

&

&

(43)

As for the synthetic approach, the extended mechanization offers the possibility to detect

defective sensors and to estimate noise terms during the integration (see Figure 10). That

said, on the contrary to the synthetic mechanization, systematic errors can be modeled and

estimated for each sensor. However, this approach requires the modification of the

GPS/INS software to accommodate the new form of mechanization equations.

33

IMU 1IMU 1

IMU 2IMU 2

IMU nIMU n

GPSGPS

NAV solNAV solNAV procNAV proc

Figure 10: Principle of extended IMU mechanization

2.5.4. Geometrically-Constrained Mechanization

As for the extended mechanization, the geometrically-constrained mechanization allows to

estimate the individual sensor errors. In this approach, multiple navigation solutions are

computed (one for each IMU) and compared at regular time intervals (Figure 11).

Unfortunately, this approach increases the computational effort and important

modifications of the GPS/INS software are required. In principle, defective sensors cannot

be detected and realistic noise terms cannot be estimated using this approach.

GPSGPS

IMU 1IMU 1

IMU 2IMU 2

IMU nIMU n

NAV sol 1NAV sol 1

NAV procNAV proc

NAV procNAV proc

NAV procNAV proc

NAV sol 2NAV sol 2

NAV sol nNAV sol n

ConstraintConstraint

Figure 11 : Principle of geometrically-constrained IMU mechanization

Consider that two IMUs are employed. Both units are integrated using the standard IMU

mechanization. The relative orientation parameters (relative orientation 2

1

bbR and lever arm

34

2

1

bba ) can be modelled and estimated as random constants, supposing their direct

determination is not accurate enough [1, 9].

0

0

2

1

2

1

=

=bb

bb

a

R

&

&

(44)

At predefined stages, the following relationships can be imposed:

( )( ) 2

1

111

2

12

2

2

11

21

21

bbw

bbib

bni

nn

bb

nb

nn

nb

bb

nb

abVV

aRrr

RRR

⋅Ω+Ω+Ω−=

⋅−=

⋅=

(45)

where Ω is the skew-symmetric form of the misalignment angles.

2.5.5. Algorithm Selection

Each of the three approaches described earlier have their advantages and disadvantages.

The synthetic mechanization requires small additional computational efforts (compared to

a single IMU mechanization) and does not necessitate any modification of the standard

GPS/INS algorithm. However, unlike the extended method, this approach does not allow

the feedback of sensor errors which might yield less optimal navigation performance. On

the other hand, [8] has shown that the MEMS-IMU biases are relatively stable for the short

trajectories encountered in some sports (e.g. lap, downhill). Furthermore, the same

research has shown that the simplified error model was suitable for the considered

application and sensors.

The geometrically-constrained approach represents an interesting option for system

calibration if the relative sensor geometry is insufficiently known. As mentioned, the

computational effort is increased considerably compared to the first two approaches. In

addition, it is more sensitive to sensor failures because errors can only be noticed at the

update stage and the measurement faults can generally not be isolated. Finally, the

geometrically-constrained approach is particularly interesting in situations where several

navigation solutions of different element of a dynamic object are matter of interest (e.g.

different parts of an athlete’s body). However, this is beyond the scope of this research. In

the subsequent sections, the synthetic IMU approach will be compared to the approach

based on extended IMU mechanization.

35

3. IMU Architectures

3.1. Introduction

This chapter aims to identify optimal sensor architecture for redundant IMUs. First, will be

considered optimal configurations of two sensors in a two dimensional space and then of

three sensors in three dimensions to illustrate the methods applied in this project.

Thereafter, optimal configurations using four sensors will be described and compared to

the work of [12]. Finally, the case of n sensor triads will be analyzed.

3.2. Assessment in Two Dimensions

This first section will verify the assumption that the best configuration for two sensors (of

the same precision) in 2D is an orthogonal configuration and that the orientation of the first

sensor is not important. Figure 12 presents the situation.

x

y

α

β

Figure 12 : Configuration of two sensors in planimetry

The design matrix H (see section 2.2) derived from this system is:

( ) ( )( ) ( )

++=

βαβααα

sincos

sincosH

Thus, the information content can be computed as defined by equation (6), that is,

Sensor 1

Sensor 2

36

( ) ( ))cos(21det 21 β⋅−⋅=⋅= HHTJ

The maximization of J yields to:

:maxJℜ∈α

( ) Ζ∈⋅+=⇒=⋅=∂∂

kkJ

,02sin 2 πβββ

π α∀ with 1max == JJ

Thus, configurations are optimal as long as 2πβ = .

3.3. Optimal Configuration with Three Sensors in Three Dimensions

Consider now the case where three sensors (with the same error characteristics) are to

placed in an optimal configuration, in a three dimensional space. Intuitively, this is

achieved when each device is placed orthogonally to each other, regardless of the

orientation of the first sensor. This is verified by the information filter approach. Note that

the partial redundancy method cannot be use here (nor in the previous section) because the

total redundancy of the system is null.

y

z

x

β

α

Figure 13: Configuration of three sensors in a three dimensional space

Figure 13 illustrates the system presented above which yields the H matrix:

( ) ( )( ) ( ) ( ) ( ) ( )

⋅⋅=

ββαβααα

sincossincoscos

0sincos

001

H

37

From this, the information volume can be computed:

( ) ( ) ( ) ( )βαβα 22 sinsindet, +=⋅= HH TJ

The maximization of J yields to:

⇒ℜ∈

Jβα ,

max 2πβα == with 1max == JJ

Hence, this result indicates the system’s optimality. Figure 14 shows the relationship

between J , α and β .

0

1

2

3

01

23

0

0.2

0.4

0.6

0.8

1

αααα [rad]ββββ [rad]

J [-

]

Figure 14: Influence of α and β on the information volume

3.4. Optimal Configuration with Four Sensors in Three Dimensions

This section considers four sensors and evaluates the optimality of configurations proposed

in the literature. Indeed, the work of [10, 12, 26] suggests three possible (optimal)

configurations:

• One sensor placed along the central cone axis while the three remaining sensors

are positioned equally around the cone of half-angle α (Figure 1)

• Equally spaced sensors around a half-angle cone α (Figure 1)

• Tetrad configuration (Figure 15).

The maximum information volume of the system is (equation (8)):

38

3704.23

43

max =

=J

This will be used to assess whether or not a configuration is optimal.

Consider the first cone configuration (Figure 1), the system’s design matrix is the

following:

( ) ( )( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( )

⋅⋅⋅⋅

=

⋅⋅

⋅⋅

αααααααα

ππ

ππ

cossinsincossin

cossinsincossin

cos0sin

100

34

34

32

32

H

The results using this information approach confirm the results of [12] and show that the

configuration is optimal, indeed,

( ) ( )( )αα 2449 cos31sin ⋅+⋅⋅=J

⇒ℜ∈

Jαmax α =70.4738° for maxJJ =

The same conclusion is also reached with the partial redundancies method. Indeed,

( )( )( ) ( )( ) ( )( ) ( )( )

T

z

⋅+⋅⋅+⋅⋅+⋅⋅+⋅=

ααααα

2222

2

cos313

1

cos313

1

cos313

1

cos31

cos3

( ) ( )( )( )[ ]( )⇒−ℜ∈

2min ααα ii zEzE α =70.4738° for 4 ,3 ,2 ,1 25.0 =≈ izi

with ( ) ( )( )( )[ ]( ) ⇒≈− 02αα ii zEzE optimal configuration.

The property related to redundancy described in equation (11) can be verified with this

example:

( )( )( ) ( )( ) runz

ii =−==

⋅++

⋅+⋅=∑

=

1cos31

1

cos31

cos322

24

1 ααα

Overall, the two methods provide the same results. The information filter allows to

determine the optimal configuration, whereas the partial redundancies approach quantifies

the amount of control for each measurement.

Consider now the second cone configuration (Figure 1), the design matrix becomes:

39

( ) ( )( ) ( )

( ) ( )( ) ( )

−−

=

αααααααα

cossin

cossin0

cos0sin

cossin0

o

H

The results provided by the information filter approach, again, confirm the results obtained

by [12] and yield an optimal configuration of

( ) ( )αα 24 cossin16 ⋅⋅=J

⇒ℜ∈

Jαmax °= 7175.54α with maxJJ =

The partial redundancies method provides the same results with 25.0≈iz for 4 ,3 ,2 ,1=i

and thus, ( ) ( )( )( )[ ] ⇒≈

− 02αα ii zEzE optimal configuration.

Finally, the tetrad configuration is now considered (see Figure 15).

y

z

x

α

β

Figure 15: Tetrad configuration proposed by A. Pejsa in [12]

The design matrix associated to this system is:

( ) ( )( ) ( )( ) ( )

( ) ( )

−−

=

0cossin

0cossin

cos0sin

cos0sin

ββββ

αααα

H

40

The results of the information filter method are:

( ) ( ) ( ) ( )( ) ( ) ( )βαβαβα 2222 coscossinsin8det, ⋅⋅+⋅=⋅= HHTJ

⇒ℜ∈

Jβα ,

max == βα 35.323° for maxJJ =

With the partial redundancies, the result also demonstrate the optimality for

== βα 35.323°, confirming the results of [12], which yields to 25.0≈iz for

4 ,3 ,2 ,1=i . Figure 16 shows the relationship between J , α and β .

0

0.5

1

1.50

0.5

1

1.5

0

1

2

3

αααα [rad]ββββ [rad]

J [-

]

Figure 16 : Influence of α and β on the information volume

Hence, the configurations proposed in the literature, are all equivalent in terms of

optimality.

3.5. Optimal Configuration with n Sensor Triads

The optimal configuration with n sensor triads in a three dimensional space is now

explained. Intuitively, skew-redundant configurations have a higher information volume

and are thus preferable. Indeed, this strategy was adopted in [9, 10]. However, as far as

sensor triads are employed, orientation between triads is unimportant with respect to the

system optimality. Consider a situation of n sensors randomly oriented. The associated

design matrix is expressed as:

41

=

),(λ

),(λ

),(λ

nn ϕ

ϕϕ

R

R

R

HM

22

11

where ),(λ ii ϕR correspond to the product of two rotation matrices (as defined in equation

(39)) and can describe any rotation in a three dimensional space. From the H matrix, the

information volume can be computed, that is:

( ) ( )),(λ),(λ),(λ),(λHHJ nnT

nnTT ϕϕϕϕ RRRR ⋅++⋅=⋅= K1111detdet

For rotation matrices, we have: 3IRR =⋅ iTi [27]. As ),(λ ii ϕR are defined as the product

of two of rotation matrices (i.e. 32 RRR ⋅=),(λ ii ϕ ), it implies that :

33223 IRRRRRR =⋅⋅⋅=⋅ TTii

Tii ),(λ),(λ ϕϕ

Note that this relationship is demonstrated in Appendix E. Hence, J becomes:

( ) ( ) 3333 detdet nnJ =⋅=++= III K

The maximum information volume can be computed as:

33

max 3

3n

nJ =

⋅=

In this way, the configuration is optimal and the information volume is independent of

ii ,λ ϕ (i.e. orientation of the triads).

The same result can be obtained using the partial redundancies method. For this system, Z

matrix can be expressed as:

42

( )( )

⋅−=

⋅⋅

⋅⋅−=

⋅⋅−=

⋅⋅⋅−=

311

113

3

11

111111

3

13

3

1

1

IRR

RRI

I

RRRR

RRRR

I

HHI

HHHHIZ

L

MOM

L

L

MOM

L

),(λ),(λ

),(λ),(λ

n

),(λ),(λ),(λ),(λ

),(λ),(λ),(λ),(λ

n

Tnn

nnT

n

nnT

nnT

nn

nnTT

n

Tnn

TTn

ϕϕ

ϕϕ

ϕϕϕϕ

ϕϕϕϕ

implying that:

]3,1[ 1

nin

nzi ⋅∈∀−=

All iz are equal, thus, ( ) ( )( )( )[ ]( ) 02 =− αα ii zEzE . The configurations are therefore

optimal, regardless of the orientation of the triads (this is also proved using simulation in

section 5.8).

3.6. Note on the Impact of Sensor Failures

In the previous section, we demonstrated that the orientation of sensor triads is irrelevant to

a system’s optimality. We now consider whether the failure of one of the sensor could alter

this conclusion and therefore render the triads’ relative orientation important. To answer

this question, consider the system presented in Figure 17.

y

z

x

y

z

x

Figure 17 : Six sensors are placed in two different configurations. The symbol indicates the sensors that will fail (superscript *).

43

The design matrices associated are:

=

3

31 I

IH and

=

)(λ ϕ,3

2 R

IH

First, the two systems yield an equivalent information volume and have equal partial

redundancies:

( ) 82det 321 =⋅== IJJ and ]6;1[ 5.02,1, ∈∀== izz ii

It was shown previously that the configurations are optimal, indeed,

836

3

max21 =

=== JJJ

Assume now that the sensors marked by in Figure 17 fail. Then, the design matrices

then becomes (the superscript * refers to system after the failures):

=

3

*1 100

010

IH and

=

)(λ ϕ,100

010*2

RH

These failures modify the optimality of the system, as well as the information volume,

leading to the following result:

630.43

54

3*max

*2

*1 ≈

=≠== JJJ

Hence, although the configurations are theoretically equivalent in the information approach

(because *2

*1 JJ = ), they are not in reality. This can be can be demonstrated using the

partial redundancy approach, where:

224.0]][[] 0 [ 2*2

12

12

12

1*1

11≈−⇒= zEzEz

( ) ( )( )( ) ( ) ( ) ( )( )( ) ( )( )

−⋅−⋅+⋅−⋅−⋅= λλϕλλϕ 222222*2 cos1

21

cos1coscos21

1coscos121

21

21

z

44

Figure 18 shows the relationship between the standard deviation of *2z and ϕ λ, . This

figure shows that 0.224 (i.e. the standard deviation of *1z ) is close to the maximum value

and that the preferred situations (i.e. skewed configuration, see Figure 18) exists. Thus,

skewed configurations are preferable in the case of sensor failures.

0.05

0.1

0.15

0.2

0.25

θθθθ [rad]λλλλ [rad]

E[(

z-E

(z)2 )]

[-]

0ππππ/4

ππππ/23⋅⋅⋅⋅ππππ/4

ππππ

0ππππ/4

ππππ/23⋅⋅⋅⋅ππππ/4

ππππ

(E[z

-E[z

]2 ])1/

2

Figure 18: Influence of λ and ϕ on the standard deviation of *2z

3.7. Conclusions

To summarize the results obtained in this chapter, we recall the main features:

• The partial redundancies and the information filter approaches give, in most cases,

similar results. That said, the partial redundancies approach is a better indicator of

the system’s optimality in certain cases.

• The best geometrical configuration of three sensors in a three dimensional space is

orthogonal to each other (i.e. sensor triads).

• The relative orientation of n sensor triads is irrelevant as long as no sensor fails,

in such case, skewed configurations are preferable.

45

4. Experimental Assessment of Redundant IMUs

4.1. Introduction

The objectives of this chapter are to assess experimentally the advantages of using multiple

IMUs. Firstly, we will evaluate the improvements in navigation (i.e. position, velocity,

orientation) resulting from the use of redundant IMUs (compared to a single IMU).

Secondly, the noise reduction stemming from the redundancy of measurements will be

experimentally assessed. Then, the algorithm of noise estimation during the processing

will be tested. Finally, the results of FDI, using the parity space method, will be exposed.

4.2. Experimental System Setup

The results presented here are based on an experiment realized with the scan2map system

(see Figure 19) which was augmented by a regular tetrahedron consisting of 4 Xsens MT-i

IMUs (see left part of Figure 19). We decided to utilize this practical skew-redundant

configuration which optimized the controllability of measurements in case of sensor failure

after the result obtained in the third chapter (see section 3.6). In order to investigate the

performance of the multi-IMU system, it was fixed rigidly to a reference system consisting

of a tactical-grade IMU (LN200) and a differential, dual-frequency GPS receiver (Javad

Legacy) [1, 28]. The system was installed as shown in Figure 19.

Tetrahedron with4 Xsens MT-i Tactical grade

IMU (LN200)

Base station

GPS antennas(on the top)

Figure 19 : (from left to right) (1) Skew-redundant IMUs placed in a tetrahedron. (2) Stéphane Guerrier holding the scan2map system. (3) Installation of the base station by Yannick Stebler

46

4.3. Improvement in Navigation Performance

This section explains the improvements brought to navigation performance when four

MEMS-IMUs (placed on a tetrahedron) are utilized (Figure 19). Table 1 summarizes the

navigation performance of the particular GPS/MEMS - Skew redundant IMU (SRIMU)

system. The orientation accuracy of this system drops below 1 deg. The experiment

confirms, on one hand, the findings of [29] where the velocity and orientation accuracies

are not dependent on the accuracy of GPS aiding (e.g. L1 or L1/L2 differential code and

carrier-phase). On the other hand, position accuracy is largely improved using dual-

frequency GPS processing with ambiguity fixing. Because of the relatively long baseline

for single-frequency CP-DGPS, the ambiguities could not be fixed, which resulted in code-

differential positioning accuracy.

L1 L1/L2 L1 L1/l2Position [m]North 0.83 0.03 0.83 0.03East 2.40 0.05 2.40 0.04Down 0.81 0.08 0.81 0.07Velocity [m/s]North 0.07 0.05 0.07 0.03East 0.07 0.07 0.07 0.05Down 0.11 0.10 0.12 0.06Attitude [deg]Roll 0.69 1.09 1.04 0.83Pitch 0.79 1.05 0.92 0.86Heading 0.42 0.68 0.67 0.62

Synthetic Extended

Table 1 : Average absolute accuracy of the tested GPS/MEMS-SRIMU system

Table 2 presents the performance improvement of the GPS/MEMS-SRIMU system,

compared to the average accuracy results obtained with the single MEMS-IMUs. The

synthetic IMU approach shows an average improvement of 30%. The extended

mechanization performs slightly better than the synthetic approach, providing for an

average improvement of 46%. This difference can be explained by the estimation of the

individual biases and the FDI scheme that run parallel to the filter (rather than in cascade

as in the synthetic IMU approach). The navigation performance is, however, not improved

by 100% as could be expected from the noise reduction (see section 4.4.1). Indeed,

residual correlations between the inertial measurements, as well as the correlations

between the filter states, most likely limit the progression of accuracy [1].

47

Table 2 : Average and maximum errors of performance resulting from the use of the MEMS-

SRIMU compared to the performance of the individually integrated MEMS-IMU sensors.

Using this technique, maximum errors are diminished, as shown in Table 2. That is, for the

synthetic IMU approach, it is of 61%, whereas for the extended mechanization maximum

errors are further reduced to a factor of two. Figure 20 illustrates how orientation error

peaks in the single GPS/MEMS-IMU integration are smoothed out by the extended

mechanization.

Figure 20 : Orientation errors after integration of a single sensor compared to extended mechanization (four sensors)

Table 2, moreover, shows a reduction in position and velocity errors, which are explained

by the significant improvement achieved in the determination of accelerations (see Figure

21). Indeed, because less noise is integrated, the computation of positions and velocities is

more precise and is smoothed out, compared to a single GPS/MEMS-IMU integration.

Synthetic Extended Synthetic ExtendedPositionNorth -29% -35% -71% -94%East -41% -51% -77% -94%Down -5% -19% -27% -82%VelocityNorth -37% -61% -82% -95%East -50% -67% -81% -93%Down -7% -44% -25% -74%AttitudeRoll -57% -67% -80% -87%Pitch -28% -41% -59% -80%Heading -21% -27% -45% -69%

RMS Maximum Error

48

Figure 21 : Position, velocity, and acceleration errors after integration of a single sensor compared to extended mechanization (four sensors)

4.4. Noise Reduction and Estimation

4.4.1. Experimental Noise Reduction

In the second chapter, we explained with equation (13) that the theoretical noise reduction

depends on a number of independent measures (i.e. the number of sensors). Therefore, the

use of four identical (MEMS-) IMUs, as in this experiment, implies a theoretical noise

reduction of a factor of two, compared to an individual (MEMS-) IMU. Hence, the

expected noise reduction is of 100%.

This theoretical noise reduction was verified by comparing the differences between the

MEMS-IMUs measurements and their best estimates (combining the measurements) with

the reference measurements provided by a tactical-grade IMU (LN200). However, this was

only realized for the gyroscopes, indeed, the (LN200) accelerometers are highly subject to

quantization noise. Thereafter, a parametric compensation was performed to remove

system errors. The remaining differences were thus considered to be only white noise. The

average noise of the four MEMS-IMUs gyros was estimated at 0.0194 [rad/s], whereas the

noise level of their best estimate at 0.0101 [rad/s]. Hence, the experimental noise reduction

is approximately 92 % which confirms the validity of the theoretical model. Figure 22

illustrates these results graphically.

49

49 49.05 49.1 49.15 49.2 49.25 49.30.9

0.92

0.94

0.96

0.98

1

1.02

1.04

1.06

1.08

1.1

Time [s]

Ang

ular

Rot

atio

n ra

te [r

ad/s

]

LN200Best estimate IMU (σ

x = 0.010 [rad/s])

MTi-G 11 (σ1 = 0.017 [rad/s])

MTi 12 (σ2 = 0.018 [rad/s])

MTi 13 (σ3 = 0.018 [rad/s])

MTi 14 (σ4 = 0.025 [rad/s])

Figure 22 : Comparison of the angular rate measurements of 4 MEMS-IMU to the reference measurements from a tactical-grade IMU

4.4.2. Assessment of the Noise Estimation Algorithm

As mentioned in section 2.3, the use of multiple IMUs allows to estimate noise level

during the processing. Figure 23 shows noise variations for a selected portion of the

experiment.

20 25 30 350

0.01

0.02

0.03

0.04

0.05

0.06

Time [s]

σσ σσA

ngul

ar r

otat

ion

rate

[rad

/s]

20 25 30 350

0.1

0.2

0.3

0.4

0.5

Time [s]

σσ σσS

peci

fic fo

rce

[m/s

2 ]

Averaging time: 0.01 [s]Averaging time: 0.1 [s]Averaging time: 1 [s]Averaging time: 10 [s]

Figure 23: Estimation of the noise level for a selected portion of the experiment

The “real” noise level was estimated using the data of the tactical-grade IMU (LN200). The

results are presented in Figure 24.

50

25 25.5 26 26.5 27 27.5 28 28.5 29 29.5 300

0.01

0.02

0.03

0.04

0.05

0.06

Time [s]

σσ σσA

ngul

ar r

otat

ion

rate

[rad

/s]

39 39.5 40 40.5 41 41.5 42 42.5 43 43.5 44

0

0.1

0.2

0.3

Time [s]

σσ σσS

peci

fic fo

rce

[m/s

2 ]

Averaging time: 0.01 [s]Averaging time: 0.1 [s]Averaging time: 1 [s]Averaging time: 10 [s]Noise level estimated with ln200

Averaging time: 0.01 [s]Averaging time: 0.1 [s]Averaging time: 1 [s]Averaging time: 10 [s]Noise level estimated with ln200

Figure 24: Comparison between the noise estimates and the “real” noise (based on ln200’s data) calculated during the experiment

Hence, the implemented adaptive filter shows results that are relatively close to those

obtained with the tactical-grade IMU approach. It allows to visualize the variation in noise

level during the processing and, consequently, to adapt the stochastic model of the EKF. In

order to test if this algorithm can improve navigation performances, we compared the

results of the GPS/INS integration with and without the noise estimation algorithm

presented above. The results are shown in Figure 25.

13 13.5 14 14.5 151

1.25

1.5

Hea

ding

err

or [d

eg]

Time [s]

with noise estimationwithout noise estimation

Figure 25 : Orientation error after integration of four sensors (extended mechanization), with and without the noise estimation algorithm

51

Orientation errors are slightly reduced, approximately 3%, because the algorithm provides

a more realistic stochastic model for the GPS/INS software. We caution, however, that this

conclusion was obtained with the results of a single experiment (presented in section 4.1),

for which the parameters influencing noise level (i.e. vibrations or other external

environmental) were considered to be constant. Nevertheless, orientation improvement

might be bigger when the noise level is expected to vary.

4.5. Fault Detection and Isolation

This section presents the results obtained with the parity space method introduced in

section 2.4. This method is based on the assumption of strictly Gaussian white noise errors.

However, MEMS-IMUs are known to be subject to systematic errors. As a result, we first

verified the white noise assumption, then, we experimentally assessed the performances of

this method with respect to detection and isolation.

4.5.1. Assessment of the Noise Characteristics of in MEMS-IMUs

This section aims to highlight if colored noise is present in Xsens MT-i (MEMS-IMUs)

measurements. Results were obtained after rotating tactical grade IMU (LN200)

measurements, which can be assumed to be perfect in this context, into each Xsens MT-i

frame and then by comparing the measurements.

-100 -50 0 50 100-1

0

1

2

Sensor 11 (Gyro): rx

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 100-2

0

2

4

Sensor 11 (Gyro): ry

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 1000

2

4

6

Sensor 11 (Gyro): rz

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 100-2

0

2

4

Sensor 12 (Gyro): rx

Time [s]Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 100-2

0

2

4

Sensor 12 (Gyro): ry

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 100-2

0

2

4

Sensor 12 (Gyro): rz

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 100-2

0

2

4

Sensor 13 (Gyro): rx

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 100-2

0

2

4

Sensor 13 (Gyro): ry

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 100-2

0

2

4

Sensor 13 (Gyro): rz

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 1000

5

Sensor 14 (Gyro): rx

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 1000

5

10

Sensor 14 (Gyro): ry

Time [s]

Aut

ocor

rela

tion

[rad

2 /s2 ]

-100 -50 0 50 1000

5

10

Sensor 14 (Gyro): rz

Time [s]Aut

ocor

rela

tion

[rad

2 /s2 ]

Figure 26: Autocorrelation of errors of the Xsens MT-i’s gyros

52

Thereafter, we analyzed the “nature” of these differences using autocorrelation techniques

(Figure 26). This clearly shows that a large portion of the non-faulty noise measurements is

composed of colored noise. Indeed, the autocorrelations below suggest that the errors are

mainly composed of white noise and of random walks.

This experiment also enabled to assess the normality of data. Figure 27 presents a “normal

probability plot” of the sensors errors. The data are plotted against a theoretical normal

distribution in such way that the normally distributed points follow a straight line. This

Figure shows quasi-Gaussian results [23].

To conclude, our simulations showed that the assumption of strictly Gaussian white noise

errors in the parity space approach is not valid in the case of MEMS-IMUs. Nevertheless,

we believe that, in sports applications, this FDI approach could be a valuable additional

feature to the system to detect gross errors and enhanced the navigation performance.

-0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08

0.001

0.003

0.01

0.02

0.05

0.10

0.25

0.50

0.75

0.90

0.95

0.98

0.99

0.997

0.999

Difference vectors for the MTi 12 ry [rad/s]

Pro

babi

lity

Figure 27: Graphical assessment of the normality of Xsens MT-i errors (sensor 12, gyro ry)

4.5.2. Performance of the Parity Space Method with MEMS-IMUs

The performance and the applicability of the parity space method to MEMS-IMUs was

also assessed using the experiment presented in section 4.2. The tactical grade IMU

(LN200) measurements (assumed to be perfect) were compared to the four Xsens MT-i’s

measurements. Thus, twelve different vectors were created (showing a quasi-Gaussian

behavior, Figure 26). Their respective standard deviations were computed and all

differences exceeding 3.5 - their standard deviation - were considered to originate from

erroneous measurements. This approach resulted in a list of “true” errors which allows

knowing, at each sample time, if one (or more) fault occurs, and to identify the defective

53

sensor. The algorithm is schematized in Figure 28. It should be noted that the assessment

of the applicability of the parity space method was only realized with gyros since the

accelerometers are highly subject to quantization noise [30].

Xsens MT-i

(Sensor 11)

Xsens MT-i

(Sensor 12)

Xsens MT-i

(Sensor 13)

Xsens MT-i

(Sensor 14)

ln200

1dσ

2dσ

3dσ

4dσ

Difference(t) > 3.5 1dσ

Difference(t) > 3.5

Difference(t) > 3.5

Difference(t) > 3.5

2dσ

3dσ

4dσ

Fault

Fault

Fault

Fault

NO Fault

NO Fault

NO Fault

NO Fault

Figure 28: Schematized algorithm for the detection and isolation of errors

The decision variable D is computed as follows:

whereff ⋅= TD

( )( ) zAAAAIf l ⋅⋅⋅⋅−= − Twwww

1

where 12=l is the number of measurements, 3=n is the number of independent

parameters, wA the matrix which converts the state space into the measurement space (as

defined in equation (38)) and z represents the vector of measurements.

The threshold T is computed according to equation (28). Two thresholds were computed

with different probabilities of false alarm FAP (i.e. 5% and 1%):

[ ]( ) [ ]22205.0 /srad 0155.0Hzrad/s/ 01.0,9,05.0 ===−= nFA nlPT σ

[ ]( ) [ ]22201.0 /srad 0198.0Hzrad/s/ 01.0,9,01.0 ===−= nFA nlPT σ

Firstly, these thresholds were compared to an optimized threshold which minimizes the

sum of false alarm probability and of misdetection [24]. The result was estimated to be

54

[ ]22/srad 0156.0ˆ =T . Thus, we observe that the theoretical threshold values computed with

the parity space method and the empirical best threshold value are of the same order of

magnitude. This experimentally derived T value can be used for future data sets. Figure

29 represents the values of D and T , and shows when fault and correct measurements

are detected successfully, as well as the occurrence of misdetections and false alarms.

46 48 50 52 54 56 58 60 62 64 66

0.005

0.01

0.015

0.02

0.025

Time [s]

D &

T [r

ad2 /s

2 ]

Decision variable (D)Threshold (T)Successful detection of non faulty measurement (95.4 %)Successful detection of faulty measurement (1.0 %)False alarm (0.4 %)Miss detection (3.2 %)

Figure 29: D values compare to T

Moreover, this figure shows the difficulty of finding a good threshold value. Indeed, even

with the best possible T value, approximately 76% of faults are not detected. The number

of false alarms is also important; the false-alarm-to-fault ratio is roughly 35%. The

performance of this test is ( ) %4.961 =+− MDFA PP . Figure 30 shows the Standford plot

of these results.

55

0 1 2 3 4 5 60

0.005

0.01

0.015

0.02

0.025

Quantile

Dec

isio

n V

aria

ble

D [r

ad2 s

- 2]

0

1

2

3

4

5

6

7

95.4%

False alarms0.4%

Misdetections3.2%

T

Successfullydetected errors

1%

3.5 σerror

Point density (per square)

Figure 30: Standford plot with TT ˆ=

Regarding the performance of the identification algorithm, Figure 31 shows that 74.4% of

faults are correctly isolated and 25.6% are not. The results developed here for T are also

valid for 05.0T .

46 48 50 52 54 56 58 60 62 64 660

0.005

0.01

0.015

0.02

0.025

0.03

Tim e [s]

D &

T [r

ad2 /s

2 ]

Decision variable (D)Threshold (T)Successful detection of non faulty measurement (95. 4 % )Successful detection of faulty measurement (1.0 % )False alarm (0.4 % )Miss detection (3.2 % )Correct isolation of fault (74.4 % )Wrong isolation of fault (25.6 % )

Figure 31: D values compared to T value with identification of performance

56

Considering the second threshold 01.0T , Figure 32 shows the values of D and 01.0T as well

as the overall performance of the test, which is ( ) %8.951 =+− MDFA PP .

45 50 55 60 65 70 75 800

0.005

0.01

0.015

0.02

0.025

0.03

Time [s]

D &

T [r

ad2 /s

2 ]

Decision variable (D)Threshold (T)Successful detection of non faulty measurement (95.8 %)Successful detection of faulty measurement (0.1 %)False alarm (0.01 %)Miss detection (4.1 %)

Figure 32: D values compared to 01.0T

Figure 33 shows the Standford plot for 01.0T .

Number of measurements

0 1 2 3 4 5 60

0.005

0.01

0.015

0.02

0.025

Quantile

Dec

isio

n V

aria

ble

D [r

ad2 s

- 2]

0

1

2

3

4

5

6

7

T

Misdetections4.1%

95.8%

False alarms0.01%

Successfullydetected errors

0.1%

3.5 σerror

Point density (per square)

Figure 33: Standford plot with 01.0TT =

57

Finally, the results of the identification algorithm are given in Figure 34.

45 50 55 60 65 70 75 800

0.005

0.01

0.015

0.02

0.025

0.03

Time [s]

D &

T [r

ad2 /s

2 ]

Decision variable (D)Threshold (T)Successful detection of non faulty measurement (95. 8 % )Successful detection of faulty measurement (0.1 % )False alarm (0.01 % )Miss detection (4.1 % )Correct isolation of fault (89.9 % )Wrong isolation of fault (11.1 % )

Figure 34: D values compared to T values with identification of performance

Thus, 05.0T is very similar to the best empirical threshold T , which indicates that 05.0T is

a more appropriate value than 01.0T . The performance of 05.0T is also higher than the

performance of 01.0T . That said, the main purpose of the FDI in non-safety critical

applications is to increase the navigation performance and to prevent gross errors. This

improvement being linked to the noise reduction, we corrected the detected faults and

assessed if this would enhance the noise reduction. Thus, for all faults detected (with 05.0T

and 01.0T ) a corrected IMU was computed based equation (46) and the erroneous

measurement was corrected.

( )( )( )( )

= ∑≠ij

z

y

xTb

bbbcorrectedi

tr

tr

tr

t

j

j

j

jiRRIMU 3

1 , (46)

Next, the noise reduction of the two sets of corrected data was studied. Without

corrections, the noise is reduced by 92% when four IMUs are used instead of one (see

section 4.4.1). With the corrected sets - at 05.0T - noise reduction was estimated at 88%,

therefore, decreasing navigation performances compared to non corrected data. However,

58

with 01.0T the noise reduction was roughly 93%, resulting in a slight increase in the

navigation performances. Consequently, 01.0T is preferred to 05.0T .

4.6. Conclusions

In the first part of this chapter, we showed that the navigation performance can be

improved by 30-50% when using four MEMS-IMUs, while maximum errors can be

reduced by a factor of two. Two integration approaches have been investigated, namely,

the synthetic and the extended mechanization. The second appears to be more optimal for

the system calibration, because the characteristic errors of individual sensors are addressed

separately. However, in the extended mechanization, where the state vector is increased by

individual biases for each sensor, estimation of biases does not reflect the reality. This

might be explained by the reduced observability of the system [1].

In the second part, we demonstrated that the expected and the experimental noise

reductions, respectively 100% and 92%, correspond. The noise estimation algorithm

additionally appears to be able to efficiently estimate the variations in noise levels. This

algorithm tends to improve slightly navigation performance when expected noise level is

constant (e.g. no vibration increases during the processing). However, it might provide

larger improvements when the noise level is expected to vary.

In the last part of this chapter, we demonstrated that the parity space method can be

successfully applied to MEMS-IMUs, because it can increase navigation performance and

prevent gross errors. It was shown that the theoretical threshold matched the best possible

threshold value, derived with reference data. However, the percentages of undetected

errors (76%), as well as the level of false alarms, remain high, showing that more complex

FDI models need to be developed. The low performance of these results may be explained

by the fact that MEMS-IMUs are largely subject to systematic errors (comparatively to

higher grade IMUs). Indeed, when dealing with MEMS-IMUs, the assumption of strictly

Gaussian white noise errors is incorrect.

59

5. Emulation of Redundant IMU Setups

5.1. Introduction

This chapter aims to present the simulation tool developed for MEMS-IMU measurement

in this project. In the context of redundant IMUs, simulation, or more precisely emulation

(which generates data corresponding precisely to the real dynamic studied), can also be

extremely helpful. Indeed, it is laborious and expensive to experimentally assess the

performance of various IMU configurations. However, this can be achieved through

simulation. In order to obtain realistic results, the following approach was followed. First,

a reference trajectory, based on a tactical-grade IMU (LN200) and a differential dual-

frequency GPS receiver (Javad Legacy), was used to generate (i.e. inverse strapdown) the

assumed perfect measurements. Then, perturbations reflecting characteristic error of IMUs

were added. These results enabled to easily emulate different configurations.

5.2. Inverse Strapdown

This section aims to define the equations that emulate the (assumed) perfect measurements

based on a reference trajectory. These equations are given in the standard notation and are

taken from [7] which presents their derivations, that is:

( ) ( ) ( ) ( )( )nb

nb

knn

ennie

nnkk

bnk

b

bbb

ib t

hRh

vwwgaRf

uw

⋅=

×+⋅+−⋅=

∆=

2

r

(47)

where the navigation frame is the local-level frame (NED) and is abbreviated by the index

n , the body frame is indicated by b , the inertial frame by i , whereas the Earth fixed

Earth centered (EFEC) frame is indexed by e.

60

5.3. Error Model for Inertial Sensors

All gyroscopes and accelerometers are subject to errors which limit the accuracy at which

angular rotations or specific forces can be measured. Detailed information on sensors

errors can be found in [15]. The error model presented here is largely based on the work

presented in [31, 32].

The basic noise model can be expressed as:

( )( ) ffff

wwww

WNBfSFMf

WNBwSFMw

++⋅⋅=

++⋅⋅=ˆ

ˆ (48)

where fw,M represent the skew-symmetric misalignment matrices, the fw,SF are the

scale factors, w and f represent respectively the realistic gyros and accelerometers

measures, w and f the idealistic measures, the fw,B are the bias vectors and fw,WN are

the white noise vectors.

The skew-symmetric misalignement matrices fw,M are defined as:

( ) ( )( ) ( )

( ) ( )

−−

−=

1

1

1

wxwy

wxwz

wywz

w

yx

zx

zy

M and

( ) ( )( ) ( )

( ) ( )

−−

−=

1

1

1

fxfy

fxfz

fyfz

f

yx

zx

zy

M (49)

where yx , zx , xy , zy , xz , yz are the misalignments between axis.

The scale factors fw,SF are defined as:

RWf

GMf

RCf

Cff

RWw

GMw

RCw

Cww

SSSS1SF

SSSS1SF

++++=

++++= (50)

where Cfw,S are the constant components of the scale factors, RC

fw,S are random constants

vectors defined by the stochastic processes: ( )fwfwRC

fw G ,,, ,σµ=S , G being a standard

Gaussian function. GMfw,S are first-order Gauss-Markov stochastic processes defined as:

61

( ) ( ) ( )ttt kGMw

GMw wSS +⋅= β& , with initial condition: ( ) 0lim 0

0

=−∞→

tGMw

tS (51)

where T1=β is the inverse of correlation time, kw is a zero-mean Gaussian white-noise

process. RWfw,S are random walk stochastic processes defined by:

( ) ( )tt kRWw wS =& , with initial condition: ( ) ( )00

RWw

RWw t SS = (52)

The bias B is defined as:

BIf

RWf

GMf

RCf

Cff

idealGSw

BIw

RWw

GMw

RCw

Cww

BBBBBB

fBBBBBBB

++++=

⋅+++++= (53)

where Cfw,B are the constant components of the bias, RC

fw,B are random constants stochastic

processes, GMfw,B are first-order Gauss-Markov stochastic processes and RW

fw,B are random

walk stochastic processes (defined as previously for the scale factors). BIfw,B are the bias

instabilities which we decided to add to original model developed by [31, 32] to modeled

more precisely the noise characteristics observed in section 5.6.2. We defined it as:

( ) ( ) ( )( )

−=⋅

= 1

0,mod

,, otherwiset

Ttiftσt BI

fw

BIkBIBIfw

w,f

B

wB

with the initial condition: ( ) ( )00, =⋅== tσt kBIBI

fw w,fwB

(54)

Finally, GSwB is the gravity sensitivity parameter.

The white noise processes fw,WN are defined as:

( ) ( )tt kfwfw wWN ⋅= ,, σ (55)

where fw,σ is the standard deviations of the white noise processes.

62

5.4. Variance Analysis

The analysis of the noise altering the output signal of random oscillators is a major issue in

various fields. These different types of noise are generally distinguished and each of them

has distinct properties. Indeed, the origins of these noises are often linked with the

oscillator environment (e.g. temperatures changes or vibrations) but they may also be

internal (e.g. thermal noise). Several variance techniques have been developed for the

analysis of these perturbations such as the Allan variance, the Hadamard variance or the

total variance [33-36]. The simplest of these methods is the Allan variance. This method

was successfully applied to the modeling of the inertial sensor errors in [30, 37-39].

The Allan variance is a method where the root mean square random-drift error is

represented as a function of the averaging time. It was invented in 1966 by David Allan

and was originally employed to study the stability of oscillators [40]. In 1998, the IEEE

standard introduced this technique as a noise identification method [41]. This method can

be used to determine the characteristics of the underlying random processes that perturb

data. The Allan variance considers five basic noise processes which can be expressed in

the appropriate notation for inertial-sensors, namely: quantization noise, white noise, bias

intensity, random walk and rate ramp. Moreover, a first-order Gauss-Markov process as

well as a sinusoidal noise can also be identified [30, 37, 42]. The Allan variance is here

used as a tool for modeling of inertial sensor errors.

Assume a data set composed of N consecutive samples, each having an interval time of

0t . From this set, n groups (or cluster) of consecutive data can be formed (with

2Nn < ). For each possible value of n , an averaging time T is associated, which is

equal to otnT ⋅= . The cluster average is defined as:

( ) ( )∫+

⋅=Tt

tk

k

k

dttT

T ΩΩ1

(56)

where ( )tΩ is the instantaneous output rate for the considered inertial sensor and ( )TkΩ

represents the cluster average that starts from the thk element of the data set and which

includes n elements. The average of the next cluster is defined as:

( ) ( )∫++

+

⋅=Tt

tnext

k

k

dttT

T1

1

1ΩΩ (57)

where Ttt kk +=+1 .

63

The Allan variance calculates the variance of the difference of two subsequent cluster

averages. Thus, it is defined as [43]:

( ) ( ) ( )[ ]22

21

TTT knextAV ΩΩσ −= (58)

The brackets of equation (58) represent the averaging operation over the ensemble of

groups. This equation can be rewritten as:

( ) ( ) ( ) ( )( )∑−

=

−⋅−⋅

=nN

kknextAV TT

nNT

2

1

22

22

1ΩΩσ (59)

Figure 35 represents schematically the computational steps of the Allan Variance.

tk+1 tk+2 tk+3 tk+n1 t0

0tnT ⋅=

tk+1+1 tk+1+2 tk+1+3 tk+1+n

Ttt kk +=+1

N

2Nn <

( ) ( )∫+

⋅=Tt

tk

k

k

dttT

T ΩΩ1 ( ) ( )∫

++

+

⋅=Tt

tnext

k

k

dttT

T1

1

1ΩΩ

( ) ( ) ( ) ( )( )∑−

=

−⋅−⋅

=nN

kknextAV TT

nNT

2

1

22

22

1ΩΩσ

Figure 35 : Schematic representation of the Allan variance

As mentioned earlier, the Allan variance measures the stability of the sensor output.

Consequently, it has to be related to statistical properties of the intrinsic random processes,

which affects sensors’ performance. There is a unique relationship between ( )TAV2σ and

the power spectral density ( )fΩS of the intrinsic random processes [43]. This relationship

is the following:

( ) ( ) ( )( )∫ ⋅

⋅⋅⋅⋅⋅⋅= Ω+∞→

t

tAV df

Tf

TffT

0 22 sin

lim4π

πSσ (60)

Its derivation is given in Appendix F.

64

5.5. Evaluation of the Noise Characteristics of Inertial Sensors

The Allan variance is an effective tool to identify different types of noise in random

processes. Indeed, equation (60) allows each noise process to be characterized by a specific

Power Spectral Density (PSD) function. This enables to estimate the contribution (and

parameters) of each noise altering a signal. Five basic noise processes will be considered

here, namely: quantization noise, white noise, bias instability, random walk and rate ramp.

Note that a first-order Gauss-Markov process and a sinusoidal noise can also be identified

([30] for more details).

5.5.1. Quantization Noise

The difference between the real analog value and the encoded digital value is called

quantization error. This error is due to the bit resolution of the analog-to-digital converter,

which is either round or truncated [44].

The angle PSD of such process is given in [45]:

( ) ( )( ) Z

zz

z

zzz T

fQTTf

TfQTf

⋅<⋅≈

⋅⋅⋅⋅⋅⋅=

2

1 ,

sin 22

22

ππ

θS (61)

where zQ is the quantization noise coefficient and zT the sample interval.

The rate PSD is related to the angle PSD through equations [30, 43]:

( ) ( ) ( )fSff ⋅⋅⋅⋅⋅=⋅⋅Ω πππ θ 222 2S (62)

and is defined as:

( ) ( ) ( )z

zzzz

z

TfQTfTf

T

Qf

⋅<⋅⋅⋅⋅≈⋅⋅⋅⋅=Ω 2

1 , 2sin

4 2222

ππS (63)

Substituting this result into equation (60) and performing the integration yields,

( )T

QT zAV

3⋅=σ (64)

65

This equation indicates that the quantization is characterized by a slope of -1 in a

logarithmical scale of ( )TAVσ . The quantization noise coefficient can be read off the

slope line at 3=T .

5.5.2. White Noise

This random process is characterized by a flat PSD which means that every frequency is of

equivalent importance in the process. Thus, this PSD is defined as [43]:

( ) 2WNf σ=ΩS (65)

Substituting equation (65) into equation (60) and performing the integration yields to:

( )T

T WNAV

σ=σ (66)

Equation (66) indicates that a white noise random process has a -1/2 slope on a logarithmic

scale. The value of WNσ can be estimated at 1=T . Figure 36 presents an example of the

Allan variance of a white noise process and the estimation of its magnitude.

10-2

10-1

100

101

102

103

104

10-10

10-9

10-8

10-7

10-6

10-5

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[-]

Original Signal (σsignal

= 4.998⋅10-6)

Best fitted straigth line (σestimated

= 4.996⋅10-6 )

Simulated with estimated parameter

Slope = -1/2

Figure 36 : Allan variance of a white noise process

66

5.5.3. Bias Instability

The bias instability originates from electronic or other sensors’ components susceptible to

random flickering [30]. The rate PSD associated with the noise is given by [43]:

( )

>

⋅=Ω

0

0

2

0

2

ff

fff

BI

πσ

S (67)

where BIσ is the bias instability coefficient.

Substituting equation (60) and integrating it yields to (details in [30]):

( ) ( )0

1

2ln2

fTT BIBIBIAV >>⋅⋅= σ

πσ (68)

Thus, at BIT , AVσ reaches the asymptotic value of ( )

2ln2

BIσπ

⋅⋅ which allows the

estimation of BIσ . Such behavior, however, may be overshadowed by the influence of

other noise terms [43]. Figure 37 and Figure 38 are examples of how bias instabilities can

be detected and how its parameters can be estimated using the Allan variance.

23 23.1 23.2 23.3 23.4 23.5 23.6-0.1

-0.08

-0.06

-0.04

-0.02

0

0.02

0.04

0.06

0.08

0.1

Time [min]

[rad

/s]

Bias Instability (+ white noise), update every 1 [s], σBI

= 3⋅10-3

Bias Instability (+ white noise), update every 10 [s], σBI

= 3⋅10-3

Figure 37 : Example of two series of “bias instabilities” and white noise

67

Figure 38 shows the Allan variance of these two series. Note that the parameters

(i.e. BIσ and BIT ) of this noise can be efficiently found with the Allan variance.

10-2

10-1

100

101

102

103

104

10-4

10-3

10-2

10-1

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad

/s]

Allan Variance of the first time serie

Allan Variance of the seconde time serie

Theoretical values of flat region of the allan variance

3⋅⋅⋅⋅10-3 ⋅⋅⋅⋅[2 ⋅⋅⋅⋅ln(2)/ ππππ]1/2 = 1.99⋅⋅⋅⋅10-2

T1 ≈≈≈≈ 0.9 [s] T2 ≈≈≈≈ 10.5 [s]

Slope = 0

Theoretical values of flat region of the allan variance

Allan Variance of the seconde time serie

10-2

10-1

100

101

102

103

104

10-4

10-3

10-2

10-1

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad

/s]

Allan Variance of the first time serie

Allan Variance of the seconde time serie

Theoretical values of flat region of the allan variance

3⋅⋅⋅⋅10-3 ⋅⋅⋅⋅[2 ⋅⋅⋅⋅ln(2)/ ππππ]1/2 = 1.99⋅⋅⋅⋅10-2

T1 ≈≈≈≈ 0.9 [s] T2 ≈≈≈≈ 10.5 [s]

Slope = 0

Theoretical values of flat region of the allan variance

Allan Variance of the seconde time serie

Figure 38 : Allan variance of the series defined in Figure 37

5.5.4. Random Walk

Random walk noise can be characterized by a trajectory consisting of taking successive

random steps. This process is controlled by the differential equation defined in the

equation (52).

In terms of PSD, a random walk is characterized by [43]:

( )2

2

⋅⋅=Ω f

f RW

πσ

S (69)

where RWσ corresponds to the random walk parameter.

Substituting equation (69) into equation (60) and performing the integration yields to:

( )3

TT RWAV ⋅= σσ (70)

From this equation, we can conclude that such random processes have a 1/2 slope on a

logarithmic scale and that the magnitude of noise RWσ can be read off a fitted straight line

at T = 3. This is illustrated in Figure 39.

68

10-2

10-1

100

101

102

103

104

10-6

10-5

10-4

10-3

10-2

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad/

s]

Original Signal (σsignal = 5⋅10-6)

Best fitted straigth line (σestimated

= 4.993⋅10-6)

Simulated with estimated parameter

Slope = +1/2

Figure 39 : Allan variance of a random walk process

5.5.5. Rate Ramp

Unlike the other types of noises described earlier, the rate ramp is not a random process. It

is defined as [43]:

( ) tCt RRk ⋅=Ω (71)

implying that:

( )2

TRTAV

⋅=σ (72)

This equation is derived in Appendix G.

Equation (72) indicates that the Allan variance of the drift rate ramp noise has a slope of 1

in a logarithmical scale. Additionally, the amplitude of this noise can be estimated at

2=T .

To conclude, Figure 40 synthesizes the elements presented in this section and shows a

schematic visual representation of Allan variance’s results.

69

Averaging Time

Roo

t Alla

n V

aria

nce

σσσσBI/(2⋅⋅⋅⋅ln(2)/ ππππ)1/2

TBI

Qz

σσσσWN

σσσσRW

CRR

21/2 31/2 31

Quantization Noise

Slope = -1 White NoiseSlope = -1/2Bias Instability

Slope = 0

Random Walk

Slope = +1/2

Rate Ramp

Slope = +1/2

Figure 40 : Schematic sample representation of Allan variance using analysis results

5.5.6. Estimation of the Quality of the Allan Variance

With real data, gradual transitions would exist between the different Allan standard

deviation slopes. Likewise, a certain amount of noise would exist due to the uncertainty of

the measured Allan variance. Indeed, the estimation of the Allan variance is computed

based on a limited number of independent clusters, hence, as this number increases, the

confidence of the estimation improves [37].

In order to estimate this confidence, the parameter δ is defined as the percentage error in

the calculation of the Allan standard deviation of the cluster, which is due to the finiteness

of the clusters:

( ) ( )( )T

TMT

AV

AVAV

σ

σσ −=

,δ (73)

where ( )MTAV ,σ denotes the estimation of the root of the Allan variance obtained with

M independent clusters. ( )MTAV ,σ converges to ( )TAVσ as the number of clusters

increases. The percentage of error is equal to [37]:

( )

−⋅=

12

1

n

Nδσ

(74)

70

where N is the total number of samples of the entire trail and n is the number of samples

contained in a cluster.

Since the size of the clusters increases along with T , the results of the Allan variance

deteriorates as the averaging time increases. This behavior can be observed in Figure 41.

10-2

10-1

100

101

102

103

104

10-4

10-3

10-2

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad

/s]

σAV

AV incertitude

Figure 41 : Typical Allan variances of MEMS gyroscope (computed with three hours static data); the spotted lines represent the standard deviation of the Allan variance values.

Equation (74) shows that ( )δσ decreases as the total number of samples N increases.

Figure 42 shows the relationship between ( )δσ , N and the averaging time which

determines n .

0 2 4 6 8 10 12 14 16 18 20 220

10

20

30

40

50

60

70

Averaging time [min]

σσ σσer

ror o

f Alla

n V

aria

nce

[%]

Error with 1h of measuresError with 2h of measuresError with 6h of measuresError with 12h of measuresError with 24h of measuresError with 48h of measures

Figure 42 : Influence of the number of sample and the averaging time to the precision of the Allan variance

71

5.6. Estimation of the Parameters of the Noise Model

In this section, we estimate the parameters of the error model, described in section 5.3,

with different approaches. In the first one, the constant part of the bias and of the scale

factor will be estimated using data collected in the experiment described in section 4.2. In

the second approach, the stochastic component of the bias, the bias instability and the

white noise will be approximated using the Allan variance method. Note that some

parameters of the error model are difficult or impossible to determine with the collected

data (i.e. the misalignment matrix and the stochastic part of the scale factor), therefore,

they are neglected.

5.6.1. Estimation of the Scale Factor and of the Constant Part of the Bias

In order to estimate the scale and bias (constant components), we considered the following

model:

( )( ) wwLNMEMS

ffLNMEMS

BSF

BSF

++⋅=

++⋅=

1

1

200

200

ww

ff (75)

where MEMSf and MEMSw correspond to the MEMS’ measurements, 200LNf and 200LNw to

the LN200’s measurements, while wfSF , represents the constant component of the scale

factor and fwB , the constant part of the bias.

A parametric compensation revealed the parameters (i.e. wfSF , and fwB , ) to be

statistically none significant. This result is confirmed by the work of [46].

5.6.2. Estimation of the Stochastic Part of the Bias and of the White Noise Parameter

The Allan variance was used to determine the stochastic component of the bias, the bias

instability and the white noise. We conducted three hours static experiment, during which

static measures of the considered sensor (i.e. Xsens MT-i) where recorded, in order to

provide inputs to the Allan variance. We consider this experiment period long enough to

give accurate results (see section 5.5.6 and Figure 41). This experiment was repeated three

times in various situations, since we assumed the noise to be sensitive to environmental

conditions (e.g. temperature, pressure). Because results may vary from one experiment to

72

another, we consider that it would be unproductive to determine precisely the noise

parameter solely based on a specific experiment. The experimental conditions in our

experiment were the following:

• Temperature ≈ 20°C, no exposure to the sun

• Temperature ≈ 16°C, no exposure to the sun

• Temperature ≈ 24°C, exposure to the sun

First, we computed the Allan variances for the gyroscopes in various environmental

conditions. We group these results in a single graphic (Figure 43).

10-2

10-1

100

101

102

103

104

10-4

10-3

10-2

10-1

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad

/s]

σAV,r

x

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,r

x

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,r

x

, exp. conditions ≈ 24 [deg], sun exposure

σAV,r

y

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,r

y

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,r

y

, exp. conditions ≈ 24 [deg], sun exposure

σAV,rz

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,r

z

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,r

z

, exp. conditions ≈ 24 [deg], sun exposure

10-2

10-1

100

101

102

103

104

10-4

10-3

10-2

10-1

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad

/s]

σAV,r

x

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,r

x

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,r

x

, exp. conditions ≈ 24 [deg], sun exposure

σAV,r

y

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,r

y

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,r

y

, exp. conditions ≈ 24 [deg], sun exposure

σAV,rz

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,r

z

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,r

z

, exp. conditions ≈ 24 [deg], sun exposure

Figure 43 : Allan variance results for the gyroscopes (spotted lines indicate the standard deviation of the measurements)

This figure shows that all sensors have a similar behavior and that the gyroscopes are

scarcely affected by variations in the environment. As a result, we can determine noise

parameters using only one experimental condition and average the different sensors’

results. To verify this estimated average, we first determined the noise parameters, and

then we created a series with the same noise parameter. We, finally, computed the Allan

variance of the simulated series and compared the results obtained with the sensor values.

The outcome of this simulation is presented in Figure 44.

73

10-2

10-1

100

101

102

103

104

10-6

10-5

10-4

10-3

10-2

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad

/s]

σσσσ WN = 0.0075

σσσσ RW = 1.904 ⋅⋅⋅⋅10-6

σσσσ BI =2.44 ⋅⋅⋅⋅10-4/(2 ⋅⋅⋅⋅ln(2)/ ππππ)1/2 = 3.67 ⋅⋅⋅⋅10-4

T ≈≈≈≈ 60 [s]

σAV,rx

σAV,rx

σAV,ry

σAV, rx,y,z

, estimated

Figure 44: Estimation of noise parameters for the gyroscopes (spotted lines indicate the standard deviation of the measurements)

Figure 45 reveals that the Allan variance of the simulated series (in black) is similar to

experimentally obtained variances, which leads us to conclude that our estimation reflects

the reality.

With respect to the accelerometers, we grouped the results obtained with the Allan

variances for different environmental conditions in a single graphic (Figure 45).

10-2

10-1

100

101

102

103

104

105

10-4

10-3

10-2

10-1

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[m/s

2 ]

σσσσAV, fz

, exp. conditions ≈≈≈≈ 20 [deg], no sun exposure

σσσσAV, fz

, exp. conditions ≈≈≈≈ 16 [deg], no sun exposure

σσσσAV, fx

, exp. conditions ≈≈≈≈ 24 [deg], sun exposure

σσσσAV, fx

, exp. conditions ≈≈≈≈ 16 [deg], no sun exposure

σσσσAV, fx

, exp. conditions ≈≈≈≈ 20 [deg], no sun exposure

σσσσAV, fy

, exp. conditions ≈≈≈≈ 20 [deg], no sun exposure

σσσσAV, fy

, exp. conditions ≈≈≈≈ 20 [deg], no sun exposure

σσσσAV, fy

, exp. conditions ≈≈≈≈ 24 [deg], sun exposure

σσσσAV, fz

, exp. conditions ≈≈≈≈ 24 [deg], sun exposure

16 [deg], no sun exposure

Figure 45 : Allan variance results for the accelerometers (spotted lines indicate the standard deviation of the measurements)

74

Figure 45 clearly shows that the accelerometers are affected by environmental variations

and that each sensor has a different behavior. As a result, we estimate each sensor

separately.

10-2

10-1

100

101

102

103

104

10-5

10-4

10-3

10-2

10-1

Averaging T ime [s]Averaging T ime [s]Averaging T ime [s]Averaging T ime [s]

Roo

t Alla

n V

aria

nce

[m/s

2 ]

σσσσ RW = 7.1498 10-5

σσσσ BI = 0.0149/(2 ⋅⋅⋅⋅ln(2)/ ππππ)1/2 =

0.0224

T ≈≈≈≈ 0.6 [s]

σσσσ WN = 0.0719

σAV,f

x

, exp. conditions ≈ 24 [deg], sun exposure

σAV,fx

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,f

x

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,f

x,estimated

Figure 46 : Estimation of noise parameters for the accelerometers (axis x) (spotted lines indicate the standard deviation of the measurements)

10-2

10-1

100

101

102

103

104

10-5

10-4

10-3

10-2

10-1

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[m/s

2 ]

σσσσ RW = 1.28 ⋅⋅⋅⋅10-5

T ≈≈≈≈ 9 [s]

σσσσ WN = 0.0132

σσσσ BI = 8.24 ⋅⋅⋅⋅10-4/(2 ⋅⋅⋅⋅ ln(2)/ ππππ)1/2 = 1.24 ⋅⋅⋅⋅10-3

σAV,fy

, exp. conditions ≈ 24 [deg], sun exposure

σAV,f

y

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,fy

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,fy,estimated

Figure 47 : Estimation of noise parameters for the accelerometers (axis y) (spotted lines indicate the standard deviation of the measurements)

75

10-2

10-1

100

101

102

103

104

10-5

10-4

10-3

10-2

10-1

Averaging Time [s]

Roo

t Alla

n V

aria

nce

[rad

/s]

σσσσ WN = 0.0214

σσσσ BI = 1.31 ⋅⋅⋅⋅10-3/(2 ⋅⋅⋅⋅ln(2)/ ππππ)1/2 = 1.96 ⋅⋅⋅⋅10-3

T ≈≈≈≈ 35 [s]

σAV,f

z

, exp. conditions ≈ 24 [deg], sun exposure

σAV,f

z

, exp. conditions ≈ 16 [deg], no sun exposure

σAV,f

z

, exp. conditions ≈ 20 [deg], no sun exposure

σAV,f

z,estimated

Figure 48 : Estimation of noise parameters for the accelerometers (axis z) (spotted lines indicate the standard deviation of the measurements)

Table 3 summarizes the results obtained in this section and compares them with the noise

characteristics provided by the constructor (i.e. Xsens) [47]. These appear to be relatively

optimistic compared to our results.

White Noise Random Walk

σWN TBI [s] σBI σRW

Gyroscopes [deg/s/ √Hz]Gyros X, Y and Z 4.3·10-2 ± 2.9·10-5 60 2.1·10-3 ± 1.1·10-4 1.1·10-5 ± 1.7·10-6

Estimated by Xsens 1.0·10-2 [ - ] [ - ] [ - ]Accelerometer [m/s 2/√Hz]Accl X 7.2·10-3 ± 4.9·10-6 0.6 2.2·10-3 ± 1.2·10-4 7.5·10-6 ± 1.1·10-6

Accl Y 1.3·10-3 ± 9.0·10-7 9 1.2·10-4 ± 3.1·10-5 1.3·10-6 ± 2.0·10-7

Accl Z 2.1·10-3 ± 1.5·10-6 35 2.0·10-4 ± 7.9·10-5 [ - ]Estimated by Xsens 0.002 [ - ] [ - ] [ - ]

Bias Instability

Table 3 : Noise estimation for Xsens MT-i using Allan variance

76

5.7. Validation of the Model

To verify the model described in the previous sections, we emulated the measurements of a

MEMS-IMU from a reference trajectory (Figure 49).

1.73451.7445

1.75451.7645

x 104

-6.4871

-6.4671

-6.4471

x 10

1700

1750

N [m]E [m]

H [m

]

x 104

x 104

Figure 49 : Reference trajectory of an alpine skier (in Plain-Joux, Switzerland,

experiment realized by Adrian Wägli)

We, then, compared the GPS/INS integration results with the emulated IMU measurements

and with a real MEMS-IMU using the same GPS data. The results are shown in Figures

50, 51 and 52.

0 5 10 15 20 25 30-0.05

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Time [s]

Pos

ition

err

or [m

]

North error with simulated dataEast error with simulated dataDown error with simulated dataNorth error with "real" dataEast error with "real" dataDown error with "real" data

Figure 50 : Comparison of the position errors with the simulated and real data

77

0 2 4 6 8 10 12 14 16 18-0.15

-0.1

-0.05

0

0.05

0.1

0.15

0.2

Time [s]

Vel

ocity

Err

or [m

/s]

VX error with simulated data

VY error with simulated data

VZ error with simulated data

VX error with "real" data

VY error with "real" data

VZ error with "real" data

Figure 51 : Comparison of the velocity errors with the simulated and real data

0 1 2 3 4 5 6 7 8-0.08

-0.06

-0.04

-0.02

0

0.02

0.04

0.06

0.08

0.1

Time [s]

Orie

ntat

ion

Err

or [r

ad]

pitch error simulated dataroll error simulated dataheading error simulated datapitch error real dataroll error real dataheading error real data

Figure 52: Comparison of the orientation errors with the simulated and real data

These figures show that the magnitudes of the errors are similar for the simulated sensors

as for the real sensors. This enables us to validate the model.

5.8. Influence of the Sensors Orientation in Redundant IMU

In the previous section, we demonstrated that our model could generate MEMS-IMU

measurements. As a result, we decided to use this tool to verify if theoretical results

obtained in section 3.5 for sensor triads (i.e. unimportance of relative orientation) were

verified by simulation. In order to do so, we simulated two sets of measurements using

four sensor triads. In the first set, all sensors have the same orientation, while, in the

second set, the sensors were positioned in the form of a regular tetrahedron (similar to

78

Figure 19). Figure 53 shows the errors resulting from the sets, in terms of position, velocity

and orientation. The results of this simulation demonstrated that sensor orientation is

irrelevant and confirmed the theoretical assumptions of the third chapter.

30 31 32 33 34 35-1.5

-1

-0.5

0

0.5

1

1.5

Time [s]

Orie

ntat

ion

erro

r [de

g]

Pitch errors with 4 sensors (orthogonal configuration)

Roll errors with 4 sensors (orthogonal configuration)

Heading errors with 4 sensors (orthogonal configuration)

Pitch errors with 4 sensors (placed on a tetrahedron)

Roll errors with 4 sensors (placed on a tetrahedron)

Heading errors with 4 sensors (placed on a tetrahedron)

10 12 14 16 18 20 22 24 26 28 300

0.05

0.1

0.15

0.2

0.25

0.3

Time [s]

Pos

ition

err

or [m

]

North errors with 4 sensors (orthogonal configuration)

East errors with 4 sensors (orthogonal configuration)

Down errors with 4 sensors (orthogonal configuration)

North errors with 4 sensors (placed on a tetrahedron)

East errors with 4 sensors (placed on a tetrahedron)

Down errors with 4 sensors (placed on a tetrahedron)

45 46 47 48 49 50-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

Time [s]

Spe

ed e

rror

[m/s

]

VNorth

errors with 4 sensors (orthogonal configuration)

VEast

errors with 4 sensors (orthogonal configuration)

VDown

errors with 4 sensors (orthogonal configuration)

VNorth

errors with 4 sensors (placed on a tetrahedron)

VEast

errors with 4 sensors (placed on a tetrahedron)

VDown

errors with 4 sensors (placed on a tetrahedron)

10 12 14 16 18 20 22 24 26 28 300

0.05

0.1

0.15

0.2

0.25

0.3

Time [s]

Pos

ition

err

or [m

]

North errors with 4 sensors (orthogonal configuration)

East errors with 4 sensors (orthogonal configuration)

Down errors with 4 sensors (orthogonal configuration)

North errors with 4 sensors (placed on a tetrahedron)

East errors with 4 sensors (placed on a tetrahedron)

Down errors with 4 sensors (placed on a tetrahedron)

45 46 47 48 49 50-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

Time [s]

Spe

ed e

rror

[m/s

]

VNorth

errors with 4 sensors (orthogonal configuration)

VEast

errors with 4 sensors (orthogonal configuration)

VDown

errors with 4 sensors (orthogonal configuration)

VNorth

errors with 4 sensors (placed on a tetrahedron)

VEast

errors with 4 sensors (placed on a tetrahedron)

VDown

errors with 4 sensors (placed on a tetrahedron)

Figure 53: Position, velocity and orientations errors of two redundant (i.e. four sensors) systems placed differently in space

5.9. Navigation Performance Improvement

In this last section, we assess the navigation performance of various redundancies using

between one and ten IMUs (based on the trajectory presented in Figure 49). The position,

velocity and orientation errors for these systems are presented in Figures 54, 55 and 56

(note that the experimental results for two sensors were taken from [1]).

1 2 3 4 5 6 7 8 9 100

10

20

30

40

50

60

70

80

90

100

# of sensors [-]

Ave

rage

pos

ition

impr

ovem

ent [

%]

RMS of simulations (synthetic)

RMS of simulations (extended)

Experimental results (synthetic)

Experimental results (extended)

Tendency cuvre (synthetic)

Tendency curve (extended)

1 2 3 4 5 6 7 8 9 100

10

20

30

40

50

60

70

80

90

100

# of sensors [-]

Max

imum

pos

ition

err

or im

prov

emen

t [%

]

Maximum errors of simulations (synthetic)Maximum errors of simulations (extended)Experimental results (synthetic)Experimental results (extended)Tendency cuvre (synthetic)Tendency curve (extended)

Figure 54 : Influence of the redundancy of the position error (right) and the maximal position error (left)

79

1 2 3 4 5 6 7 8 9 100

10

20

30

40

50

60

70

80

90

100

# of sensors [-]

Max

imum

vel

ocity

err

or im

prov

emen

t [%

]

Maximum errors of simulations (synthetic)

Maximum errors of simulations (extended)

Experimental results (synthetic)

Experimental results (extended)

Tendency cuvre (synthetic)

Tendency curve (extended)

1 2 3 4 5 6 7 8 9 100

10

20

30

40

50

60

70

80

90

100

# of sensors [-]

Ave

rage

vel

ocity

impr

ovem

ent [

%]

RMS of simulations (synthetic)

RMS of simulations (extended)

Experimental results (synthetic)

Experimental results (extended)

Tendency cuvre (synthetic)

Tendency curve (extended)

Figure 55 : Influence of the redundancy of the velocity error (right) and the maximal velocity error (left)

1 2 3 4 5 6 7 8 9 100

10

20

30

40

50

60

70

80

90

100

# of sensors [-]

Ave

rage

orie

ntat

ion

impr

ovem

ent [

%]

RMS of simulations (synthetic)

RMS of simulations (extended)

Experimental results (synthetic)

Experimental results (extended)

Tendency cuvre (synthetic)

Tendency curve (extended)

1 2 3 4 5 6 7 8 9 100

10

20

30

40

50

60

70

80

90

100

# of sensors [-]

Max

imum

orie

ntat

ion

erro

r im

prov

emen

t [%

]

Maximum errors of simulations (synthetic)Maximum errors of simulations (extended)Experimental results (synthetic)Experimental results (extended)Tendency cuvre (synthetic)Tendency curve (extended)

Figure 56 : Influence of the redundancy of the orientation error (right) and the maximal orientation error (left)

As expected, these Figures show an inverse correlation between errors and the number of

sensors. Furthermore, the curve tendencies appear to have a comparable form to the

expected noise reduction shown in Figure 4. However, and as explained earlier, residual

correlations between inertial measurements, as well as correlations between filter states,

limit the reduction of navigation errors (i.e. errors in position, velocity and attitude)

compared to the noise reduction.

80

5.10. Conclusions

This chapter has shown that, with an appropriate error model, MEMS-IMUs measurements

can be successfully emulated from a reference trajectory. For that reason, we consider our

model able to simulate any sensor architecture and to assess its resulting navigation

performance. This could provide helpful information for applications that require a limit

precision by determining the needed number of sensors to fulfill the requirements.

Additionally, we verified, through simulation, the theoretical results obtained for the

orientation of the triads (section 3.5). We identified the relationship between the different

navigation performance parameters and the number of sensors employed. That said, these

results should be considered with caution as they are based on a single trajectory. Further

test would be required to define a useable relation.

81

6. Conclusion and Perspectives

6.1. Synthesis

This master project is part of a larger investigation that aims to develop a low-cost GPS/

INS system for performance analysis in sports activities. The main objective of this project

was to investigate orientation improvement when using redundant inertial sensors.

To summarize the results obtained in this research, we remind the main features:

• A new method (i.e. partial redundancy), originally applied in geodetic networks,

allows to assess the optimality of redundant system. This method has

demonstrated that, in most cases, relative sensor orientation is unimportant for the

optimality of the system, as long as no sensor fails; however, in this case, skewed

configurations are preferable.

• Redundant IMUs can improve the determination of position, velocity, acceleration

and orientation. For example, with four sensors, measurements results gain

approximately 30% in accuracy for the synthetic approach and 50% for the

extended approach.

• We demonstrated and quantified the noise reduction resulting from the

redundancy of measurements. We, then, successfully employed this redundancy to

estimate variations of the noise level during the processing. Experimental results

show that this provides a more realistic error model to the EKF and improves

navigation performance.

• This research has proved that, when applied to sports activities, the parity space

method can advantageously be applied with MEMS-IMUs. On the other hand,

percentages of undetected errors, as well as the level of false alarms, remain high

and, consequently, show a need for more complex FDI models to be developed.

• We have demonstrated in this thesis that MEMS-IMU measurements can be

emulated with an appropriated error model. This allows to theoretically estimate

navigation performance of any redundant systems. This offers a tool to design

systems according to specifications.

To conclude, the increased accuracy provided by redundant IMUs has the potential of

better bridging the gaps in GPS data. In addition, this project, in providing the opportunity

to estimate noise levels and to correct errors, demonstrated why redundant IMUs are

interesting to enhance the performance in many (sports) applications.

82

6.2. Perspectives

The conclusions of this thesis have high potential for a wide range of fields, where a single

IMU is insufficient to provide satisfactory reliability or/and accuracy (e.g. direct

georeferencing, virtual reality, military application, pedestrian navigation). The results

should, nevertheless, be considered with caution as they are based on a single experiment.

Further experimentation should be realized to verify these results.

83

7. References

[1] A. Waegli, S. Guerrier, and J. Skaloud, "Redundant MEMS-IMU integrated with GPS for Performance Assessment in Sports," in IEEE/ION Position Location and Navigation Symposium, Monterey, 2008.

[2] A. Waegli, B. J.-M., and J. Skaloud, "L'analyse de performance sportive à l'aide d'un système GPS/INS low-cost: évaluation de capteurs inertiels de type MEMS," Revue XYZ, vol. 113, 2007.

[3] A. Waegli, A. Schorderet, C. Prongué, and S. J., "Accurate Trajectory and Orientation of a Motorcycle derived from low-cost Satellite and Inertial Measurement Systems," in 7th ISEA 2008, Biarritz, France, 2008.

[4] A. Waegli and J. Skaloud, "Turning Point - Trajectory Analysis for Skiers," InsideGNSS, 2007.

[5] J. How and N. Pohlman, "GPS Estimation Algorithms for Precise Velocity, Slip and Race-track Position Measurements," SAE Motorsports Engineering Conference & Exhibition, 2002.

[6] K. Zhang, R. Deakin, R. Grenfell, Y. Li, J. Zhang, W.N.Cameron, and D. M. Silcock, "GNSS for sports - sailing and rowing perspectives," Journal of Global Positioning Sytems, vol. 3, pp. 280-289, 2004.

[7] A. Waegli, "Performance Assessment in Sports based on Navigation Systems," in Geodetic Engineering Laboratory. vol. phd Thesis: EPFL, 2008 (to be published).

[8] A. Waegli, J. Skaloud, P. Tomé, and J.-M. Bonnaz, "Assessment of the Integration Strategy between GPS and Body-Worn MEMS Sensors with Application to Sports," in ION GNSS, Fort Worth, Texas, 2007.

[9] I. Colomia, M. Giménez, J. J. Rosales, M. Wis, A. Gómez, and P. Miguelsanz, "Redundant IMUs for Precise Trajectory Determination," 2006.

[10] S. Sukkarieh, P. Gibbens, B. Grocholsky, K. Willis, and H. F. Durrant-Whyte, "A Low-Cost Redundant Inertial Measurement Unit for Unmanned Air Vehicles," The Internation Journal of Robotics Reasearch, 2000.

[11] A. Osman, B. Wright, S. Nassar, A. Noureldin, and N. El-Sheimy, "Multi-Sensor Inertial Navigation Systems Employing Skewed Redundant Inertial Sensors," in ION GNSS 19th International Technical Meeting of the Satellite Division,, Fort Worth, Texas, 2006.

[12] A. J. Pejsa, "Optimum Skewed Redundant Inertial Navigators," AIAA Journal, vol. 12, 1974.

[13] http://www.intensiondesigns.com/itd-biotensegrity/biotensegrity/papers/images/geometry_of_anatomy_images/fig_2_platonic_solids.jpg.

[14] H. Dupraz and M. Stahl, Théorie des Erreurs. Lausanne, 1994. [15] D. H. Titterton and J. L. Weston, Strapdown inertial navigation technology. London,

1997. [16] N. F. Zhang, "The uncertainty associated with the weighted means of measurement

data," Metrologia, vol. 43, 2006.

84

[17] K. C. Daly, E. Gai, and J. V. Harrison, "Generalized Likelihood Test for FDI in Redundant Sensor Configurations," Journal of Guidance and Control, vol. 2, 1979.

[18] E. Gai, J. V. Harrison, and K. C. Daly, "FDI Performance of Two Redundant Sensor Configurations," IEEE Transactions on Aerospace and Electronic Systems, vol. AES - 15, 1978.

[19] A. Medvedev, "Fault Detection and Isolation by Continuous Parity Space Method," Automatica, vol. 31, 1995.

[20] M. A. Sturza, "Navigation System Integrity Monitoring Using Redundant Measurements," The Journal of The Institute of Navigation, vol. 35, 1988-89.

[21] U. Krogmann, "Artificial Neural Network for Inertial Sensor Fault Diagnosis," in Symposium Gyro Technology, Stuttgart, Germany, 1995.

[22] B. W. Parkinson and P. Axelrad, "Autonomous GPS Integrity Monitoring Using the Pseudorange Residual," Journal of the Institute of Navigation, vol. 35, 1988.

[23] R. A. Johnson, Probability and Statistics for Engineers, 2005. [24] S. Feng, W. Y. Ochieng, D. Walsh, and R. Ioannides, "A Measurement Domain

Receiver Autonomous Integrity Monitoring Algorithm," GPS solution, 2006. [25] A. Gelb, Applied Optimal Estimation, 1994. [26] G. Dissanayake, S. Sukkarieh, E. Nebot, and H. Durrant-Whyte, "The Aiding of Low-

Cost Strapdown Inertial Measurement Unit Using Vehicle Model Constraints for Land Vehicle Applications," IEEE Transactions on Robotics and Automation, vol. 17, 2001.

[27] J. Skaloud and K. Legat, "Navigation Techniques," EPFL, 2007. [28] J. Skaloud, J. Vallet, K. Keller, G. Veyssière, and O. Kölbl, "An Eye for Landscapes -

Rapid Aerial Mapping with Handheld Sensors," GPS World, vol. May 2006, pp. 26-32, 2006.

[29] A. Waegli and J. Skaloud, "Assessment of GPS/MEMS-IMU Integration Performance in Ski Racing," in ENC, Geneva, Switzerland, 2007.

[30] H. Hou, "Modeling Inertial Sensors Errors using Allan Variance," in Dept. Geomatics Eng. vol. M.S. thesis: Univ. Calgary, 2004.

[31] M. E. Parés, "On the development of an IMU simulator," in Institute of Geomatics. vol. M.S. thesis: Universitat Politècnica de Catalunya, 2008.

[32] M. E. Parés, R. J.J., and C. I., "Yet another IMU Simulator: Validation and Applications," 2008.

[33] D. A. Howe, R. L. Beard, A. Greenhall, F. Vernotte, W. J. Riley, and T. K. Peppler, "Enhancements to GPS Operations and Clock Evaluations Using "Total" Hadamard Deviation," IEEE Transactions on Ultrasonics, Ferroelectrics and Requency Control, vol. 52, 2005.

[34] F. Vernotte, E. Lantz, and J. J. Gagnepain, "Oscillator Noise Analysis: Multivariance Measurement," IEEE Transactions on Instrumentation and Measurement, vol. 42, 1993.

[35] F. Vernotte, E. Lantz, J. Groslambert, and J. J. Gagnepain, "A new Multi-Variance Method for the Oscillator Noise Analysis," in IEEE Frequency Control Symposium, 1992.

[36] F. Vernotte and G. Zalamansky, "A Bayesian Method for Oscillator Stability Analysis," IEEE Transactions on Ultrasonics, Ferroelectrics and Requency Control, vol. 46, 1999.

85

[37] N. El-Sheimy, H. Hou, and X. Niu, "Analysis and Modeling of Inertial Sensors Using Allan Variance," IEEE Transactions on Instrumentation and Measurement, vol. 57, 2008.

[38] J. M. Strus, M. Kirkpatrick, and J. W. Sinko, "GPS/IMU Development of a High Accuracy Pointing System for Maneuvrering Platforms," InsideGNSS, 2008.

[39] Z. Xiang and D. Gebre-Egziabher, "Modeling and Bounding Low Cost Inertial Sensors Errors," 2008.

[40] D. W. Allan, "Statistics of Atomic Frequency Standards," in Proceedings of IEEE, 1966.

[41] I. S. 1293, " IEEE Standard Specification Format Guide and Test Procedure for Linear, Single-Axis, Non-gyroscopic Accelerometers," 1998.

[42] "IEEE Recommended Practice for Inertial Sensor Test Equipment, Instrumentation, Data Acquisition and Analysis," 2005.

[43] I. S. 952, "IEEE Standard Specification Format Guide and Test Procedure for Single –Axis Interferometric Fiber Optic Gyros.," 1997.

[44] Wikipedia, "Quantization Error," http://en.wikipedia.org/wiki/Quantization_error, Ed. [45] A. Papoulis, "Probability, Random Variables, and Stochastic Process, Third Edition,

McGraw-Hill, Inc.," 1991. [46] J.-M. Bonnaz, "Analyse du comportement de capteurs inertiels en trajectographie," in

Geodetic Engineering Laboratory. vol. Master Thesis: EPFL, 2007. [47] Xsens, "MT-i, Miniature Attitude and Heading Reference System,

http://www.xsens.com/Static/Documents/UserUpload/dl_41_leaflet_mti.pdf." [48] M. M. Tehrani, "Ring Laser Gyro Data Analysis with Cluster Sampling

Technique," in Proceedings of SPIE, 1983.

86

8. Appendix

8.1. Appendix A: Proof of equation (11)

Theorem: ( ) unztrn

ii −==∑

=1

Z

Proof: ( ) ( )( )1−⋅⋅⋅−= RHPHRZ Ttrtr

( )( )( )( )( )( )

( )( )( )( ) ( )( )( ) ( )

un

trtr

trtr

tr

tr

tr

un

TT

TT

TT

TT

−=−=

⋅⋅⋅⋅⋅−⋅=

⋅⋅⋅⋅⋅−⋅=

⋅⋅⋅⋅⋅−⋅=

⋅⋅⋅⋅⋅−=

−−−−

−−−−

−−−−

−−−

II

HRHHRHRR

HRHHRHRR

HHRHHRRR

RHHRHHR

1111

1111

1111

111

8.2. Appendix B: Proof of equation (13)

Theorem: The variance of the best estimate x is n

w xn

iiix

σσσ == ∑=1

22ˆ assuming

iix ∀= σσ

Proof: The best estimate can be expressed as ∑=

⋅=n

iii xwx

1

ˆ where ∑=

=n

iiw

1

1. The

expected value of x is then given by: ( ) ( )∑=

⋅=n

iii xEwxE

1

ˆ . Moreover, the variance of a

quantity y is defined as ( )( )[ ]22 yEyEdef

y −=σ . Thus the variance of x is:

( )

( )( ) ( )( ) ( )( ) ijxExwxExwxExwE

xEwxwE

n

i

n

jjjjiii

n

iiii

n

iiiiix

≠∀

−⋅⋅−⋅+−⋅=

⋅−⋅=

∑ ∑∑

= ==

=

,1 1

2

1

2

2

1

2ˆσ

87

Since all ix are independent, then the ( )( )ii xEx − are uncorrelated and thus all

( )( ) ( )( )[ ] 0=−⋅− jjii xExxExE [15]. Hence, 2xσ can be expressed as:

( )( )[ ] ∑∑∑===

⋅=⇒⋅=−⋅=n

iiix

n

iii

n

iiiix wwxExEw

1

22ˆ

1

22

1

222ˆ σσσσ

Since it is assumed that iix ∀= σσ then iww ix ∀= , since the iw are proportional to

the precision of the measures. This implies that ∑=

=⇒=⋅=n

ixxi n

wnww1

11 . Thus

nx

x

σσ =ˆ

8.3. Appendix C: Proof of equation (19)

Theorem: The best least square estimate x of the equation bnxHz ++⋅= is zHx ⋅= *ˆ

where ( ) TT* HHHH ⋅⋅= −1

Proof: xHzvvxHbnxHz ⋅−=⇒+⋅=++⋅= where bnv += , the best estimate

x is obtained when sum of squared differences (i.e.v ) is minimized, mathematically:

( ) ( ) ( )( )xHzxHzvv ⋅−⋅⋅−=⋅ TT minmin

The partial differentiation of vv ⋅T yields to the determination of x , indeed,

( ) ( )( ) ( ) ( )( )

( )( ) ( ) ( ) ( )( )

( )( ) ( ) ( )( ) zHHHx

zHxHHHHxzxHzH

xHzx

HxzxHzHxzx

xHzHxzx

xHzxHzx

⋅⋅⋅=⇒

⋅=⋅⋅⇔=−⋅⋅−+⋅−⋅−⇔

=⋅−∂∂⋅⋅−+⋅−⋅⋅−

∂∂⇔

=⋅−⋅⋅−∂∂⇔=⋅−⋅⋅−

∂∂

− TT

TTTTTT

TTTTTT

TTTT

ˆ0ˆˆ

0ˆˆˆˆ

0ˆˆ0ˆˆ

This implies that zHx ⋅= *ˆ where ( ) TT* HHHH ⋅⋅= −1

88

8.4. Appendix D: Proof of equation (23)

Theorem: *l HHIS ⋅−=

Proof: Firstly, [ ]TPHA =−1 has to be proved to be the correct inverse of the matrix

=

P

HA

*

In order to do so, the equality lIAA =⋅ −1 must be true.

[ ]

( ) ( )( )( ) ( )

( ) ( )( )( ) ( )

( )( )

( )( ) ( )

l

nlnnl

nlnn

nlnnl

TTn

nlnnl

TTTTT

T

T**

T*

I

I0

0I

I0

HPHHI

I0

PHHHHHHH

PPHP

PHHH

PHP

HAA

=

=

⋅⋅⋅=

⋅⋅⋅⋅⋅⋅=

⋅⋅⋅⋅

=

=⋅

−×−

−×

−×−

−×−

−−

1

11

1

Since A is a square with full rank matrix, then:

[ ] SHHPPHHP

HPHAAI +⋅=⋅+⋅=

⋅=⋅= − *T*

*T

l1

This implies that: *l HHIS ⋅−=

89

8.5. Appendix E: Proof of equation (42)

Theorem: ∑=

⋅⋅=n

iz

y

xTb

bb

i

i

i

iw

r

r

r

n 1

1Rw , the derivation is similar for bf and bm

Proof: ( ) bwll

Twwll

Tw

bww

bbw

bw

bw

bw

lRAARAlΠw ⋅⋅⋅⋅⋅=⋅= −−− 111

Assuming IR ⋅= 20σb

wbwll

, it follows:

( )[ ]( ) b

wTw

bb

Tbb

bb

Tbb

bw

Tww

Tw

b

nwnwww lARRRR

lAAAw

⋅⋅⋅++⋅=

⋅⋅⋅=−

1

1

11 L

It can shown that 3IRR =⋅ iwiw bb

Tbb , indeed,

( ) ( ) ( ) ( ) ( )( ) ( )

( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( )

( ) ( )

( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

( ) ( ) ( ) ( ) ( ) ( )( ) ( )

( ) ( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

( ) ( ) ( ) ( ) ( ) ( )( ) ( ) ( ) ( ) ( ) 1sinsincoscoscos

0cossincossincoscos

0cossinsinsincossincoscos

0cossincossincoscos

1cossin

0sincossinsincossin

0cossinsincossincoscossin

0cossinsinsincossin

1cossinsincossin

since, ,

100

010

001

sin0cos

sincoscossinsin

coscossincossin

sinsincoscoscos

0cossin

cossinsincossin

2222233

32

2231

23

2222

21

2213

12

2222211

3

333231

232221

131211

=+⋅+⋅=

=⋅⋅−⋅⋅−==⋅−⋅⋅+⋅⋅=

=⋅⋅−⋅⋅==+=

=⋅⋅+⋅⋅−==⋅−⋅⋅+⋅⋅=

=⋅⋅−⋅⋅==+⋅+⋅=

∀=

=

=

−⋅−⋅−⋅−−⋅−

−⋅−⋅−−

⋅−⋅−=⋅

λϕλϕλϕϕλϕϕλ

λλλϕλλϕλϕϕλϕϕλ

ϕϕϕϕλϕϕλ

λλϕλλϕλλϕϕλϕϕλλϕλϕλ

θλ

λλϕλϕϕλϕλϕϕλ

λϕλϕλϕϕ

λϕλϕλ

a

a

a

a

a

a

a

a

a

I

aaa

aaa

aaa

iwiw bb

Tbb RR

Thus,

[ ]( ) ∑=

⋅⋅=⋅⋅++=n

1

133

1

iz

y

xTb

bbw

Tw

b

i

i

i

iw

r

r

r

nRlAIIw L

90

8.6. Appendix F: Proof of equation (60)

Theorem: ( ) ( ) ( )( )∫ ⋅⋅⋅= Ω+∞→

t

tAV df

fT

fTfT

0 22 sin

lim4π

πSσ

Proof : This proof is based on [27, 30, 48]. Firstly, the Allan variance defined by the

equation (58) is transformed as follows:

( ) ( ) ( )[ ]

( ) ( ) ( ) ( )

( ) ( )[ ] ( ) ( )TTTT

TTTT

TTT

knextknext

knextknext

knextAV

2222

2222

22

2

1

22

12

1

ΩΩΩΩ

ΩΩΩΩ

ΩΩσ

⋅−+⋅=

⋅⋅−+=

−=

Secondly, the definition of ( )TkΩ (i.e. the equation (56)) is used to compute ( )Tk2

Ω

and ( )Tnext2

Ω , that is,

( ) ( ) ( ) ( ) tddtttT

TdttT

TTt

t

Tt

tk

Tt

tk

k

k

k

k

k

k

′⋅⋅′⋅=⇔⋅= ∫ ∫∫+ ++

,11

22 RΩΩΩ

where ( ) ( ) ( )tttt ′⋅=′ ΩΩR , is the autocovariance function. Assuming the ( )tΩ to be

stationary ( )tt ′,R can be written as ( ) ( ) ( )τRRR =−′=′ tttt, . Moreover, the basic

relationship for stationary processes between the two sided PSD is [27]:

( ) ( ) ( ) ( )∫ ∫− −

⋅⋅⋅⋅−Ω→∞

⋅⋅⋅⋅−

→∞Ω ⋅⋅=⇔⋅⋅=t

t

t

t

fi

t

fi

tdfefSdtef τπτπ ττ 22 limlim RRS

Thus, it follows that:

( ) ( )

( )

( ) ( ) ( )df

f

Tff

T

dftddtefT

tddtttT

T

t

tt

Tt

t

Tt

t

fit

tt

Tt

t

Tt

tk

k

k

k

k

k

k

k

k

⋅⋅⋅⋅⋅

=

⋅′⋅⋅⋅⋅=

′⋅⋅′⋅=

∫ ∫∫

∫ ∫

− Ω∞→

+ + ⋅⋅⋅⋅−

− Ω∞→

+ +

2

2

22

22

sinlim

1

lim1

,1

ππ

τπ

S

S

Additionally, ( )Tk2

Ω is not a function of time, thus ( ) ( )TT knext22

ΩΩ = . Then, the

term ( ) ( )TT knext22

ΩΩ ⋅ is obtained using equation (56), that is,

91

( ) ( ) ( ) ( ) tddtttT

TdttT

TTt

t

Tt

Ttk

Tt

tk

k

k

k

k

k

k

′⋅⋅′⋅=⇔⋅= ∫ ∫∫+ ⋅+

+

+ 2

22 ,

11RΩΩΩ

Substituting ( )tt ′,R , it yields:

( ) ( )

( ) ( ) ( )df

f

Tfef

T

tddtttT

T

Tfit

tt

Tt

t

Tt

Ttk

k

k

k

k

⋅⋅⋅⋅⋅⋅

=

′⋅⋅′⋅=

⋅⋅⋅⋅

− Ω→∞

+ ⋅+

+

∫ ∫2

22

2

22

sinlim

1

,1

ππ

πS

Consequently,

( ) ( ) ( )[ ] ( ) ( )

( ) ( ) ( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( ) ( ) ( )

⋅⋅⋅⋅⋅⋅⋅⋅−⋅⋅⋅⋅⋅⋅

=

⋅⋅⋅−⋅⋅⋅

=

⋅−=

⋅−+⋅=

∫∫

− Ω∞→− Ω∞→

⋅⋅⋅⋅

− Ω∞→

2

2

2

4

2

2

22

222

22222

2sinsinlim

sinlim2

1

sin1lim

1

2

1

f

TfTffi

f

Tff

T

dff

Tfef

T

TTT

TTTTT

t

tt

t

tt

Tfit

tt

knextk

knextknextAV

ππππ

ππ

π

SS

S

ΩΩΩ

ΩΩΩΩσ

Since ( )fΩS is defined as an even matrix [48] the above equation becomes:

( ) ( ) ( ) ( ) ( ) ( ) ( )

( ) ( ) ( )

( ) ( )( )2

4

0

2

4

2

2

2

2

4

22

sinlim4

sinlim

2

2sinsinlim

sinlim2

1

fT

Tff

f

Tff

T

f

TfTffi

f

Tff

TT

t

t

t

tt

t

tt

t

ttAV

⋅⋅⋅⋅⋅⋅=

⋅⋅⋅⋅⋅

=

⋅⋅⋅⋅⋅⋅⋅⋅−⋅⋅⋅⋅⋅⋅

=

∫∫

Ω∞→

− Ω∞→

− Ω∞→− Ω∞→

ππ

ππ

ππππ

S

S

SSσ

92

8.7. Appendix G: Proof of equation (72)

Theorem: ( )2

TRTAV

⋅=σ

Proof: The equation (71) states: ( ) tRt ⋅=Ω , thus Ω and nextΩ are expressed using the

equations (56) and (57), that is,

[ ]22)(1

)(1

kk

Tt

t

Tt

ttTt

T

RdttR

Tdtt

T

k

k

k

k

−+⋅=⋅⋅⋅=⋅⋅= ∫∫++

ΩΩ

[ ]222)()2(

1)(

1TtTt

T

RdttR

Tdtt

T kk

tt

t

Tt

Ttnext

k

k

k

k

+−⋅+⋅=⋅⋅⋅=⋅⋅= ∫∫++

+ΩΩ

Then using the definition of the Allan variance (i.e. equation (59)), it yields to

( )

[ ]

( )

( )

( )

2

2)2(2

)2(2

42244)2(2

)(2)2()2(2

)2(2

1)(

22

2

24

2

1

222

2

22

1

222222

2

22

1

2222

2

22

1

2

RT

nNnNT

RT

TnNT

R

ttTTttTTtnNT

R

tTtTtnNT

R

nNT

nN

k

nN

kkkkkk

nN

kkkk

nN

knextAV

⋅=

⋅−⋅⋅−⋅⋅

⋅=

⋅⋅−⋅⋅

=

+⋅⋅−⋅−⋅−⋅⋅−⋅+⋅⋅−⋅⋅

=

++⋅−⋅−⋅⋅−⋅⋅

=

−⋅⋅−⋅

=

=

=

=

=

ΩΩσ

Thus, ( )2

TRTAV

⋅=σ

93

“Aime la vérite, mais pardonne à l’erreur.”

Voltaire, Discours en vers sur l’homme, Deuxième discours de la liberté


Top Related