parametric diagnostics of flight control and propulsion for...

27
Submitted to AIAA Journal of Guidance, Control, and Dynamics June 2004 1 Parametric Diagnostics of Flight Control and Propulsion for Rocket Ascent Dimitry Gorinevsky Honeywell Labs, Fremont, CA 94539 John R. Bain Honeywell Space Systems, Houston, TX 77058 Gordon Aaseng Honeywell Space Systems, Glendale, AZ 85308 Corresponding author: Honeywell Labs, 47102 Mission Falls Court, Fremont, CA 94539; phone: (510) 360-7906, e-fax: (208) 545-6408; e-mail: [email protected], [email protected] Abstract. This paper describes a case study of model-based diagnostics system development for a space launch vehicle application. The diagnostics algorithms described in the paper can work with on-board or telemetry flight data. The key innovation is in designing a distributed diagnostics system that directly uses parametric data, such as sensors data. The computations are partitioned between the subsystems and a central IVHM system to match the system knowledge partitioning. The subsystem algorithms use detailed systems models. The central algorithms use cross-vehicle fault signature knowledge. The developed algorithms provide fault condition estimates that allow for consistent detection of incipient performance faults and abnormal conditions.

Upload: others

Post on 28-Jan-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

Submitted to AIAA Journal of Guidance, Control, and Dynamics • June 2004

1

Parametric Diagnostics of Flight Control and Propulsion for Rocket Ascent

Dimitry Gorinevsky∗ Honeywell Labs, Fremont, CA 94539

John R. Bain Honeywell Space Systems, Houston, TX 77058

Gordon Aaseng Honeywell Space Systems, Glendale, AZ 85308

∗Corresponding author: Honeywell Labs, 47102 Mission Falls Court, Fremont, CA 94539; phone: (510) 360-7906, e-fax: (208) 545-6408; e-mail: [email protected], [email protected]

Abstract. This paper describes a case study of model-based diagnostics system

development for a space launch vehicle application. The diagnostics algorithms

described in the paper can work with on-board or telemetry flight data. The key

innovation is in designing a distributed diagnostics system that directly uses

parametric data, such as sensors data. The computations are partitioned between the

subsystems and a central IVHM system to match the system knowledge partitioning.

The subsystem algorithms use detailed systems models. The central algorithms use

cross-vehicle fault signature knowledge. The developed algorithms provide fault

condition estimates that allow for consistent detection of incipient performance faults

and abnormal conditions.

Page 2: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

2

1. Introduction This paper demonstrates an application of model-based parametric diagnostics approach to a space

vehicle GN&C-related application. This is done through a case study detailed in this paper. The

need for integrated vehicle health management (IVHM) has been long recognized. IVHM

functionality is a major requirement for the development of next generation space vehicles. The

main benefits expected from an IVHM system include:

• Improving system safety by detecting faults and acting on this information. The actions can

include control system reconfiguration, alerting, or mission management.

• Reducing cost and time required for maintenance and vehicle readiness checks. This is

particularly important for the next generation reusable vehicles that are supposed to have

much shorter turnaround and mission preparation time.

Achieving these benefits requires evaluating the system health state from the vehicle sensor data.

The health state evaluation logic might reside in the vehicle avionics or be based on ground and use

telemetry or flight-logged data. A basic approach to fault detection is through a range or threshold

checking of the sensor data. Recent advances in fault detection technology extend this in two

directions. One is developing IVHM systems capable of vehicle-wide integration of the fault

detection results to achieve a more accurate understanding of the fault root cause. Another is using

parametric data from the sensors and detailed system models to discriminate between component

deterioration and environmental influences.

The most prominent example of currently deployed IVHM system for an aerospace vehicle is the

Boeing 777 CMC (Central Maintenance Computer) developed by Honeywell in the 90s [9]. The same

technology was subsequently used in Honeywell’s Primus Epic CMC and AMOSS ground based

aircraft maintenance automation systems. The CMC computer is a separate avionics piece

connected through the vehicle buses to avionics of all the vehicle subsystems. Each subsystem has

its own diagnostics function, such as BIT/BITE (Built-In Test / Built-In Test Equipment). The

subsystems provide diagnostics messages in a predefined format. The CMC collects these messages

and reasons out the fault root causes that might have caused all these messages. As one example, a

Page 3: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

3

single failed sensor or a loose connector might cause several hundred subsystems to announce

faults of different types. In this design, the subsystem diagnostics are typically developed and

implemented by the subsystem hardware vendors who have the detailed application knowledge.

The central IVHM function – the CMC – integrates these diagnostics and embeds the knowledge

how the subsystems are integrated in the vehicle.

This work discusses a step beyond the above described discrete diagnostics IVHM architecture and

towards parametric diagnostics. Instead of dealing with discrete data (fault/no fault, in/out of bound

measurement, etc), the parametric data comprising grey-scale sensor values are considered.

Compared to the discrete data diagnostics, the processing of parametric data has a number of

important benefits:

In general, parametric diagnostics are more detailed and accurate than discrete diagnostics

because they are based on more detailed data.

Parametric estimation of fault condition allows for providing more specific alerts, compared

to a mere presence of a fault, such as: ENGINE THRUST IS 5% BELOW THE NOMINAL.

Parametric estimation and diagnostics provides data necessary for prognostics trending.

This brings a potential benefit of a identifying an incipient fault condition and taking a

premeditative action (e.g., performing maintenance) before the fault has fully developed.

Parametric estimation can be used to discriminate between sensor drift and a system

problem

Typically, this work no exception, parametric diagnostics are based on using models of the system.

These models allow predicting system output variables based on the ambient conditions and input

variables. Such models are expensive and developing them requires detailed application

knowledge. Therefore parametric diagnostics, at least initially, should be focused on most critical

systems and functions, where the benefits are the strongest.

An approach to parametric fault estimation presented herein assumes that at least some of the

computations should be performed on the vehicle subsystem level. At the same time, a central

Page 4: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

4

IVHM system is necessary for integrating the subsystem processing results and providing vehicle

level estimation and prognostics. Development of parametric diagnostics for vehicle subsystems

typically requires using detailed simulation models of subsystem dynamics. On another hand,

development of the central IVHM function requires understanding of how the subsystems are

integrated within the vehicle and what are cross-system fault signatures. The goal of this work is to

demonstrate these system architecture concepts in a case study of parametric diagnostics: estimation

of fault parameters for flight control (GN&C) and propulsion systems of the vehicle. In addition to

these, flight controls actuator – main engine gimbal actuator – is included as a separate subsystem in

this study.

The state of the art in parametric diagnostics is mostly defined by the FDI (Fault Detection and

Identification) work done in the controls community over last decade. A substantial literature on the

subject has been published, including several books, e.g., see [2], also see the surveys [2, 5, 7]. This

literature assumes availability of detailed dynamical models in analytical form, typically linear

models. Detailed fault models are usually assumed to be an integral part of the dynamical models.

Various analytical algorithms have been developed that estimate presence and sometimes

parameters of the faults from the available data. These approaches have practical implementation

deficiencies for the problem in hand.

Instead of analytical linear models, most of realistic detailed simulations in aerospace industry use

comprehensive nonlinear models developed by teams of people. When used in an IVHM

development, these are best described as ‘black-box’ models. This means a software implementation

of the model can be run to obtain results, but detailed analytical structure of the model is usually

unavailable.

The published FDI methods usually assume that the model is available as a single integrated block.

At the same time, the detailed knowledge of and modeling capability for different vehicle

subsystems is usually vested with different subsystem developers. A practical design of IVHM

system should take this fact into account.

In addition to the subsystem-level detailed FDI, there is a need for a centralized insight into faults

that manifest across different subsystems. This is not addressed in the existing FDI literature.

Page 5: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

5

In this work the algorithms use linearized fault models for model prediction residuals. The

approach is an extension of the optimal fault estimation approach applied in [4]. However, instead

of point-wise estimation in [4], the faults are estimated by a linear-quadratic receding horizon

algorithm. Some of the prior work on the receding horizon estimation can be found in [1, 8].

2. Technical problem

The parametric IVHM system design approach described below has a broad applicability. In this

work it is explained and demonstrated in a case study. The object of the case study is a simulation

for ascent of a “Space Shuttle class” launch vehicle. The simulation provides a representative

environment for application of model-based Integrated Vehicle Health Management (IVHM). In the

case study, model-based IVHM algorithms attempts to determine selected “deterioration

parameters” seeded in the simulation.

The algorithms demonstrated in this paper assume that the data from flight (GN&C), propulsion,

and engine gimbal subsystems are available for processing at a high sampling rate. The algorithms

can reside in on-board avionics of the vehicle. In that case, the data is assumed to be transmitted

over on-board avionics buses of the vehicle and available to health management processors as

required. Alternatively the data can be transmitted in a telemetry stream and the algorithms reside

in an on-ground health management system. The conceptual computational logic and algorithms

partitioning are the same in either case.

The goal is to demonstrate estimation of the fault parameters describing degradation in the

propulsion engine (thrust loss), degradation or damage of the vehicle aerodynamic surface (drag

increase), degradation in the thrust vectoring system (gimbal actuator sluggishness), and signal drift

fault in a GN&C sensor (pitch sensor offset). The approach of this work consists of the following

three steps. First, a detailed simulation model is set up to simulate the relevant telemetry (or the

vehicle bus) data. The model allows simulating the faults in question. The results of the simulation

with the (known) seeded faults are logged and stored in a data file for subsequent processing. The

Page 6: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

6

data in the file is subsequently replayed and processed by the developed fault estimation

algorithms. These algorithms have no information on the seeded faults and can be validated by

comparing the seeded and estimated fault parameters.

As a second logical step in the development, we discuss subsystem-level algorithms for computing

residuals from sensor data using subsystem prediction models. These residuals indicate an off-

nominal behavior of the system and are indicative of the fault presence. The residual computation

and prediction models are described in Section 3 of this paper.

Finally, the algorithms for estimating fault parameters from the residuals are described in Section 4.

These centralized algorithms use the prediction residuals from all subsystems and could reside in a

separate central IVHM processor. An important part of the estimation algorithm is computation of

fault sensitivities (fault signatures) for each of the faults. These fault signatures are subsequently

used by a receding horizon estimation algorithm performing a multivariate curve fit to compute

updates for the fault parameter estimates.

The discussion in this paper ends wth obtaining the fault parameter estimates. Getting these

estimates in consistent and accurate way is the most technically challenging part of the parametric

health management functionality. At the same time, the largest application value is in using the

obtained estimates on the fault parameters for alerting and decision making. The alerting and

decision logic is highly dependent on the application and accepted procedures. Discussing this logic

is outside of scope of this work.

As a part of the case study, a simulation model of the launch vehicle ascent was developed. This is a

simplified model of a Space Shuttle class vehicle that is based on data published in the open

literature. Detailed engineering simulations of Space Shuttle used by NASA have in excess of 100

states, and thousands of parameters. Details of such complex simulations are understood by teams

of people. Unlike that, the simulation set up in this study has on the order of 10 states, a few dozen

parameters, and is understood by a single person. The simulation describes the rocket launch flight

phase. Though the detail of the simulation is not presented here, below are some highlights.

Page 7: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

7

A set of ordinary differential equations (ODE) numerically integrated in the simulation comes from

physics model describing dynamics of the system. The simulated ODE system is stiff, because the

timescales for rocket motion and fast GN&C actuator are very different. The magnitude of the

variables runs many orders of magnitude. For instance, the control toque is required to move a 5⋅106

lb thrust engine bell for angle tracking of ~10-3; while the final value of the achieved altitude is

~5⋅105.

We simulate a two-stage vehicle with liquid fuel (H2 and O2) delivering a “medium” payload to

LEO. We consider in-plane dynamics only for a planar equatorial flight that terminates in a 0°

inclination orbit. The modeling includes variation of mass, center of gravity, and moment of inertia

with the propellant expenditure. An aerodynamical model is borrowed from an asymmetric vehicle

(first stage drag minimum is at small positive angle of attack. The model assumes non-rotating

spherical Earth with inverse square gravitational field and exponential atmosphere. In the example

below, the first stage trajectory and vehicle data are primarily considered. This is because air drag is

still present and one of the considered fault symptoms is a drag increase

A simplified model of the GN&C system is implemented as a part of the simulation. In a trajectory

following guidance/control scheme, the engine bell is gimbaled to attain desired pitch angle while

the thrust is maintained at 100% (it is a function of altitude and stage). We assume a full-state

measurement to be available for the vehicle dynamics. A Neighboring Extremal (NE) closed-loop

guidance uses a point mass simulation to calculate the minimum fuel trajectory. NE guidance is

implemented as a full-state proportional feedback, which does not steer back to the original optimal

trajectory. Instead, at each guidance calculation, if the vehicle has deviated from the optimal path,

NE guidance provides the desired control for the optimal trajectory from that point to the desired

endpoint. A PID controller tuned using “trials-and-errors” is used for the main engine gimbal

actuator to keep the vehicle on the trajectory.

The simulation is PC based and programmed in Mathworks’ Simulink as a “continuous time”

model with Runge-Kutta integration of ODEs. The simulation covers the 370 sec trajectory of ascent

Page 8: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

8

into an 80 x 150 nm equatorial orbit. The vehicle launch mass is 1.0438⋅108 slugs; mass flow rate is

constant in each stage; staging occurs upon expending of first stage propellant at tstage = 153.54 s. The

sensor and control data is logged at a 0.1 s sampling interval through the simulation run and saved

into a data file. The simulation model includes an ability to add (seed) the faults described in

Section 4.

The next section describes some detail of the prediction models. These are based on the same

underlying dynamics knowledge as the simulation but are differently structured, towards the needs

of the IVHM algorithms. One major difference is that simulation includes closed-loop GN&C

algorithms, while the prediction model just uses the actually applied control effort without a need

to detail how it was computed.

3. Residual computation and prediction model

The three subsystems considered in this case study (GN&C, Propulsion, and Gimbal) each include

an embedded model that allows computing prediction residuals from the available sensor data. The

fault parameter estimation from these residuals is performed by the central IVHM system. This

section discusses the prediction models and residual computations performed by the subsystems,

the following section – the fault estimation computations in the central IVHM system.

The general idea of the prediction model is to divide the variables describing subsystem operation

into two parts: inputs (this includes ambient condition data and data from other subsystems) and

outputs. The prediction model allows computing the outputs given the inputs. The differences

between the predicted (computed) and actually observed outputs make the prediction residuals.

Assuming that the prediction model accurately describes the nominal subsystem behavior, the

residuals should be zero if there is no fault. Nonzero residuals indicate that a fault is present.

It is important to emphasize that the prediction models for subsystems conceptually differ from the

simulation models. While the prediction model uses some of the available data to predict other data,

the simulation model is used to predict system evolution over the time. These models might be

Page 9: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

9

based on the same system knowledge and contain some of the common blocks, but nevertheless are

conceptually very different.

Two types of data are processed in a predictive model of each subsystem. First is the sensor data

from the sensors attached to this subsystem. In addition to being processed internally, this data is

provided to other subsystem through the data buses. The second type of data is the external data

from other subsystems. This comprises both the raw sensor data and the derived variables

computed by other subsystems. The prediction residuals computed by the subsystems are provided

to the central parametric IVHM system.

A generic view of subsystem-level computations is outlined in Figure 1. One type of the processed

data is the data from the sensors attached to this subsystem. This data is provided to other

subsystems. The second type of data is the external data from other subsystems: the raw sensor data

and/or the derived variables computed by other subsystems. The model-based computations in the

subsystem provide the derived variables to other subsystems as well as the prediction residuals for

the central IVHM system.

predictionresidualsSu

bsys

tem

Internal data data Externaldata

sensors

Data from other subsystems

variableestimates

To central IVHM system

To other subsystems

Model-based

estimation

Figure 1: Member System for parametric diagnostics and prognostics.

Let us now consider the subsystem computations design for the three subsystems in this case study:

G&C, Propulsion, and Gimbal. These computations are based on the shared data available on the

vehicle avionics buses and in the telemetry stream. The data assumed to be available at a high

smapling rate include the following.

Page 10: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

10

Vehicle state data

Downrange angle κ measured in rad

Altitude h measured in ft

Velocity v measured in ft/s

Flight path angle γ measured in rad

Engine gimbal angle δ measured in rad

Engine rotational rate ωe measured in rad/s

Vehicle rotational rate ω measured in rad/s

Pitch angle θ measured in rad

In the analysis to follow, the data listed above is collected in a state vector X,

[ ]T

evhX θωωδγκ= (3.1)

Vehicle accelerations and rates

The state variables in vector X (3.1) enter the right hand sides of equations of motion describing the

simulation model. In addition to these variables, some of the derivatives in left-hand side of these

equations are assumed to be available. These include

Vertical acceleration dv/dt measured in ft/s2

Flight path angle rate dγ /dt measured in rad/s

Vehicle rotational acceleration dω/dt measured in rad/s2

The vertical acceleration dv/dt and rotational acceleration dω/dt of the vehicle are directly measured

by the accelerometers in the inertial navigation system (INS) of the vehicle or can be computed as a

transformation of such direct accelerometer measurements. The flight path angle rate can be

estimated from high rate flight path angle measurements, other attitude measurements and the

actual path measurements. In the analysis to follow, the three mentioned accelerations are collected

in a vector A

T

dtd

dtd

dtdvA ⎥⎦

⎤⎢⎣⎡= ωγ

(3.2)

Page 11: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

11

Gimbal actuator servo data

In the diagnostics algorithms to follow, the vehicle dynamics variables (3.1) and (3.2) are

complemented by the Gimbal actuators subsystem measurements. The following two variables are

assumed available:

Gimbal actuator position xact measured in rad

Gimbal actuator servo command xd measured in rad

The data from the above described sensors obtained in a simulation run is shown in Figure 2. The

time interval in the plots covers the first stage ascent. The data in the plots is the raw data and is

used by the below described algorithms to estimate the fault parameters for the vehicle.

50 100 150-1.405

-1.4

-1.395

-1.39

Downrange angle (rad)

50 100 150

5

10

15

x 104 Altitude (ft)

50 100 150

2000

4000

6000

8000 Velocity (ft/s)

50 100 150 0.40.60.8

11.2

Flight path angle (rad)

50 100 1500

5

10

15 x 10

-3 Engine gimbal angle (rad)

50 100 150 0

5

10

15

x 10-3Engine rotational rate (rad/s)

50 100 1500

5

10

15

x 10-3 Vehicle rotational rate (rad/s)

50 100 150

0.60.8

11.21.4

Pitch angle (rad)

50 100 150-0.05

0

0.05

Gimbal command (rad)

Time (s)50 100 150

-0.04-0.02

00.020.04

Gimbal position (rad)

Time (s) Figure 2: Sensor data traces obtained in the first stage ascent simulation

Page 12: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

12

3. 1 Flight dynamics (GN&C) subsystem

Consider first the residual computation for the flight dynamics (GN&C) subsystem. The prediction

model is based on the same dynamics equations as used in the simulation model. The two equations

for center of mass motion have the form

[ ]γmgDδαTm

v sin)cos(1 −−+=& (3.3)

hrγvγmgLδαT

mvγ

s ++−++= cos]cos)sin([1

& (3.4)

The two attitude dynamics equations for the in-plane motion of the main rocket body and the

engine are

Qmmδmlω

αDαLαDαLlmωδmωJ

QmmδmlTδωm

αDαLmllmmωδmωJ

e

eeee

ee

cpee

)(]sin

)cossin()sincos[(cos

)(sinsin

)sincos]()[(cos

2

**

2*

**

+−−

+−++−=+

++−+

+−+=+

&&

&&

(3.5)

The mass, rotational inertia, and geometric parameters of the vehicle such as

secpee rlllmmmJJ ,,,,,,,, *** , in (3.3)--(3.5) are assumed to be known or are computed on-line from

the vector (3.1). The parameters L and D are respectively lift and drag. These forces can be

computed from the state vector X. The lift and drag depend on the pitch angle α = θ - γ and the

Mach number (the speed v). Computing L and D requires also knowing the atmospheric density,

which can obtained through the density tables based on the attitude h. The steering torque Q in

(3.5) is a result of the thrust vectoring and can be computed as

acttorque xKQ = , (3.6)

where Ktorque can be approximately assumed to be a constant. Finally, the thrust T in (3.5) comes

from Propulsion subsystem computations (see below).

Page 13: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

13

The two coupled rotational dynamics equations (3.5) can be resolved with respect to the angular

accelerations ω& and eω& . The predictive model uses only calculated ω& . Adding the calculated ω& to

the calculated accelerations (3.3)--(3.4) leads to the predictive model equations that can be compactly

presented in vector form as

( )txTXFv

actA ,,,=⎥⎥⎥

⎢⎢⎢

ωγ&

&

&

(3.7)

Comparing the computed accelerations (3.7) against the observed accelerations (3.2) yields a 3 x 1

prediction residual vector for the GN&C system

( )txTXFAr actAflight ,,,−= (3.8)

3. 2 Propulsion subsystem

In this work, no detailed model of the propulsion was used. This was because of the limited work

scope and the model availability. No dynamical state variables of the rocket engine are modeled in

the simulation example, no sensor measurements specific to the propulsion are assumed. In a real

vehicle there are additional engine-specific measurements. A predictive model using the additional

measurements could produce additional residuals correlated with the thrust loss. This would

improve estimation accuracy.

The propulsion system closely interacts with the flight control system through the thrust it delivers.

In the simulation, a simple propulsion model is used. The thrust is assumed to be available as a

function of time and altitude. The prediction model assumes that sufficient sensor data is available

to the propulsion subsystem for continuously calculating the thrust using an embedded model. It is

further assumed that this computation is accurate in the absence of the faults. The computed thrust

is used by other subsystems.

Page 14: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

14

The simple model used for the thrust prediction was exactly the same as one used in the initial

simulation. This model can be expressed in the form

( )tXFT T ,= , (3.9)

where X is as in (3.1) and only the second component of the vector X is used in (3.9).

3. 3 Gimbal actuation subsystem

We used a simplified model for the hydraulic actuator of the main engine gimbal suspension. The

actuator is used for vectoring the main engine thrust and, thus, is the main control actuator. The

attitude torque produced by the thrust force depends on the vectoring angle actx controlled by the

actuator in accordance with (3.6). The bandwidth of the control feedback loop used to guide the

vehicle along the trajectory is critically limited by the sluggishness of this actuator. The gimbal

actuation includes a servo-system for tracking the desired vectoring angle dx and the closed-loop

feedback linearizes the dynamics of the system. The assumed linearized model of the dynamics is a

first-order model of the form

factdact ωxxx )( −=& , (3.10)

where fω is the servo-system bandwidth indicative of the actuator sluggishness. An increase on

the sluggishness reduces the bandwidth.

The predictive model is based on the actuator model (3.10) and uses the available actuator position

measurement actx and the servo-system setpoint measurement dx . The predictive model should

verify if these two measurements satisfy the system dynamics equation (3.10). This equation cannot

be verified directly because there is a derivative actx& in the l.h.s. (3.10). Differentiation is a

noncausal operation and cannot be implemented exactly. The solution is to use an approximate

smoothing differentiator. To extract the information about the sluggishness, the model prediction

residual is computed as follows

Page 15: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

15

( )dactfs

acts

gymbal xxs

xs

sr −+

−+

= ωττ

1, (3.11)

where s is the Laplace variable (differentiator operator), sτ is a smoothing filter pole (time

constant) and the two transfer functions describe the filtering operators applied to the two

respective signals. The two filters in the r.h.s. (3.11) have proper and strictly proper transfer

functions and can be easy implemented. Note that if (3.10) is satisfied then the residual in (3.11) is

exactly zero. Applying a smoothing filter is a linear transformation of the signal. It does not offset

the residual, just brings about a filtering delay while enabling differentiator implementation.

In this work, the filter pole sτ was selected by a trial and error method and was chosen to be

ss 05.0=τ . A more systematic design of differentiating filters should take into account the

statistical properties of the noise influencing the differentiated signal. Such noise is not modeled in

(3.10) but would likely influence a real-life sensor signal.

3. 4 Computed residual

The residuals are functions of time. The overall residual vector collecting the subsystem residuals

(3.8) and (3.10) has the form

⎥⎦

⎤⎢⎣

⎡=

)()(

)(trtr

trgymbal

flight (3.12)

The residual vector r (3.12) has four components. Figure 3 illustrates these four components

computed for the first stage ascent data in Figure 2. The faults have been included in the simulation

in Figure 2, therefore the prediction residuals in Figure 3 are nonzero. These residuals are further

used by the fault parameter estimation algorithm.

Page 16: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

16

0 20 40 60 80 100 120 140 160 0

0.5

1

1.5 ACCELERATION [ft/s2]

0 20 40 60 80 100 120 140 160 0

1

2

3 x 10

-3 FLIGHT ANGLE RATE [rad/s]

0 20 40 60 80 100 120 140 160 -5

0

5

10 x 10

-4 PITCH ACCELERATION [rad/s2]

0 20 40 60 80 100 120 140 160 -5

0

5 x 10

-3 SERVO RATE [rad/s]

TIME [s] Figure 3: Residuals computed from the first stage ascent simulation data.

4. Estimating fault parameters The computational data flow architecture for the parametric IVHM processing as applied to the

problem at hand is outlined in Figure 4. The overall parametric IVHM system consist of a Central

IVHM System and thee subsystems: Propulsion, GN&C, and Gimbal. Thus partitioned, the three

subsystems can be implemented on separate processors communicating through on-board

databases, or telemetry link, or telemetry link and on-ground network. The key advantage of the

architecture is that software development requiring detailed domain knowledge for a subsystem

can be relegated to the subsystem provider.

Each of the subsystems collects data from affiliated sensors and computes residuals as discussed

earlier in Section 3. This section discusses the central processing of the residuals in some detail.

Page 17: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

17

4.1 Parametric IVHM architecture

The GN&C subsystem computations are shown as the upper left block in Figure 4 schematics and

are described in Subsection 3.1. The arrow labeled “Predicted acceleration residuals” describes

computation of the acceleration residuals (3.7) in accordance with (3.8). These residuals are sent for

further processing into the Central IVHM System.

The Propulsion subsystem logic in Figure 4 (the middle block on the left) is discussed in Subsection

3.2. This logic does not produce any residuals that are specific for the engine. However it uses the

available sensor data to produce an estimate of the engine thrust (3.9) that is used by the GN&C

subsystem for computing the predicted acceleration residuals.

Prop

ulsi

on

subs

yste

m

Internal data data External data

Propulsion sensors

Propellant component flows

ThrustestimateModel-

basedestimation

Predicted acceleration residual

Internal data data External data

GN&C sensors:

Thrust vector data Airflow probe data

Model-based

estimation

Internal data data External data

Position sensors

Commanded position

EngineorientationModel-

based estimation

Actuatorrate residuals

Central Parametric HM: Estimate Fault Parameters

Sluggish Gimbal Signature

Air Drag Change Signature

Reduced Thrust Signature

Fault Signatures

Central IVHM System Subsystem HM

GN

&C

su

bsys

tem

G

imba

l su

bsys

tem

Pitch Sensor Drift

Figure 4: Conceptual organization of the fault parameter computations

Page 18: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

18

The residual computations in the Gimbal actuator subsystem are as outlined in Subsection 3.3 and

are shown as the lower left block in Figure 4. This subsystem provides a filter output (3.11) to the

Central IVHM subsystem. An arrow depicting flow of this data is labeled as “Actuator rate

residual” in Figure 4.

The Central Parametric HM block of Central IVHM Subsystem in Figure 4 is where the fault

estimates are computed. In addition to the residuals (3.12), the Central Parametric HM block uses

fault signatures as an input. These fault signatures and their computation is defined below in this

subsection.

4.2 Fault modeling

As mentioned above, four different faults are considered in this case study. The intensities of these

faults make a fault parameter vector

⎥⎥⎥⎥

⎢⎢⎢⎢

=

offsetsensor pitchsssluggishne gimbal

change dragair lossthrust

f (3.13)

The components of the vector f (3.13) are measured in percent. The fault parameters (3.13) are

introduced into the simulation and prediction models as follows.

Engine thrust

When simulating the engine thrust fault, the thrust value T computed in (3.9) and used throughout

various parts of the prediction (and simulation) models at each time instant is replaced by the value

Tf = (1 – 0.01 f1) T, (3.14)

where f1 is the percentage of thrust loss, the first component of the vector f (3.13). In the prediction

model, this fault is associated with the Propulsion subsystem. The engine thrust reduction can be

caused by multiple reasons. It might be an indication of the remaining useful life depletion for a

reusable engine.

Page 19: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

19

Air drag change

The air drag increase might be caused by a damage of the aerodynamic surface of the launch

vehicle. With an aerodynamic return vehicle being a part of the launch vehicle (such as Space

Shuttle orbiter or a planned OSP vehicle) the drag increase might indicate damage to the protective

covering of the return vehicle surface. This could be a serious problem, such as one in the Columbia

accident. This fault is simulated by replacing an instantaneous value of the air drag D in the flight

dynamics equations (3.3) and (3.5) with an increased drag

Df = (1 + 0.01 f2) D, (3.15)

where f2 is the percentage of the drag increase, the second component of the vector f (3.13). In the

prediction model, this fault is associated with the GN&C (Flight dynamics) subsystem.

Gimbal actuator sluggishness

The sluggish response of the gimbal actuator air drag increase can be caused by a hydraulic valve

problem or pressure loss in the hydraulic system of the actuator. This fault is simulated by changing

the bandwidth fω in gimbal dynamics equation (3.10) into the fault-modified value

1001808 3

0fωω f ⋅−= π

, (3.16)

where f3 is the percentage of the sluggishness increase, the third component of the vector f (3.13).

The nominal open-loop bandwidth of the actuator is assumed to be 10 35.0 −= sω (the actuator time

constant of 2.86 s). The 100% sluggishness increase of the fault corresponds to the open-loop

bandwidth reduction to 10 21.0 −= sω (the actuator time constant of 4.77 s). In the prediction model,

this fault is associated with the Gimbal actuator subsystem.

Pitch sensor fault

In this work, pitch sensor fault was modeled as a gain (calibration) error of the sensor. The pitch

sensor fault is simulated by replacing an instantaneous value of the pitch angle θ used in the GN&C

control law computations with the modified value

Page 20: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

20

θf = θ /(1 + 0.01 f4) , (3.17)

where f4 is the percentage of the sensor gain loss, the last (fourth) component of the vector f (3.13).

In the prediction model, this fault is associated with the GN&C (Flight dynamics) subsystem. Both

the prediction and simulation models of the flight dynamics include GN&C control law as an

integral, but separate, part.

4.3 Sensitivity modeling

The fault parameters in (3.14)—(3.17) enter the system dynamics in a nonlinear way. The central

IVHM algorithms used in this work are based on a fundamental assumption that the system

dynamics changes caused by these faults is relatively small in most operating conditions. Thus, it

becomes possible to linearize the dependence of the residuals on the faults without loosing too

much accuracy. The estimation algorithms we use are based on such linearization and are

demonstrated to give sufficiently accurate estimates of the faults.

Let us proceed in describing the linearized dependence of the residuals on the faults in a rigorous

way. In general, the residual data processed by the Central IVHM System and used for the

estimation is the sampled values of the residual vector r (3.12) collected over some interval of time.

This data can be presented in the form

⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢

∆+

∆+∆+

=∆

)(

)2()(

)(

)0;,,(

Ntr

trtr

tr

NtY

b

b

b

b

b

M

, (3.18)

where tb is the time of the data collection beginning, ∆ is the sampling interval, and N is the number

of the sequential samples collected. The vector Y has the size of 4N .

Page 21: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

21

The residual data (3.18) is based on the prediction models described in Section 3 and assume a

nominal system behavior, no fault. This zero intensity of fault, f = 0, is indicated by the last

argument in the l.h.s. (3.18). It is possible to extend the prediction model to include the fault models

(3.14)—(3.17) and use it to predict the system output in a case when a known fault is present. The

residual data obtained assuming a known fault will be further denoted as );,,( fNtY b ∆ .

Linearizing the residual data dependence on the fault vector f yields

fNtSNtYfNtY bbb ),,()0;,,();,,( ∆+∆≈∆ , (3.19)

where S is the 4N x 4 Jacobian of the map );,,( fNtY b ∆ with respect to its last argument

0

);,,(),,(=∂

∆∂=∆f

bb f

fNtYNtS (3.20)

The matrix S in (3.19), (3.20) is further called a sensitivity matrix. The four columns of this matrix

contain the signatures of the four faults in consideration.

The map );,,( fNtY b ∆ is not available analytically. Rather, pointwise values of this map can be

computed by running the extended prediction model (with the fault models incorporated).

Therefore the sensitivity matrix is computed by a secant (finite difference) method. This is done by

in turn incrementing each component of the fault vector and computing the respective residual data

vector );,,( fNtY b ∆ . The normalized increments of the residual vector then yield secant estimates

for the columns of matrix S. Such computation of the sensitivity matrix could be done prior to the

flight, assuming that the vehicle would closely follow the nominal trajectory. The pre-computed

fault signatures (columns of the matrix S) are stored and used for on-line processing of the flight

data stream. The on-line computational logic using these signatures is outlined in Figure 4.

There is an inaccuracy in using a precomputed sensitivity matrix since the actual vehicle state is not

exactly the same as for the planned trajectory and for the nonlinear map );,,( fNtY b ∆ the

computed matrix S depends on the actual system states. An alternative, more accurate, approach is

Page 22: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

22

to compute the fault signatures on-line, for the actual system state data. Here is how this can be

implemented.

At each point of time, in addition to the nominal model prediction residuals (3.12), four more

residuals are calculated. Denote by r(t; f) the residual calculated by assuming the fault f in the

prediction model. Then the nominal model prediction residual is

)0;()(0 == ftrtr (3.21) The four fault signatures (columns s(j) of the matrix S) can be computed by running four additional

modified copies of the prediction model alongside with the nominal prediction model within the

on-line data processing framework. All five models are receiving the same data stream from the

vehicle, but assume different faults. Let e(j) , (j =1, …,4), be unit fault vectors with all components

zero except the unit component j. Then the expressions for the signatures are as follows

dtrdetrts

jj )();()( 0

)()( −= , (3.22)

where d is in the finite difference step size. Our case study used d=1. One percent faults do not

violate the linearization assumptions. The matrix S is then computed by sampling the signature data

as follows

⎥⎥⎥⎥

⎢⎢⎢⎢

∆+∆+

∆+∆+=∆

)()(

)()()()(

),,(

)4()1(

)4()1(

)4()1(

NtsNts

tstststs

NtS

bb

bb

bb

b

L

MOM

L

L

, (3.23)

The fault sensitivities (signatures) computed by a finite difference method are illustrated in Figure 6.

The four columns in the array of plots correspond to the four faults in consideration. The rows

correspond to the four residuals. Each plot shows a part of the time dependence s(j)(t). The combined

and sampled data in each column of the Figure 5 plot array make a column of the matrix S.

Page 23: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

23

0 50 100 150 -1.5

-1

-0.5

0

ACC

ELER

ATIO

N

GIMBAL SLUGGISHNESS

0 50 100 150

-4

-2

0

x 10 -3

FLIG

HT

ANG

LE R

ATE

0 50 100 150

-2

-1

0

1 x 10

-3

PITC

H A

CC

ELE

RA

TIO

N

0 50 100 150 -2

-1

0

1

2 x 10

-4

EN

GIN

E A

NG

LE R

ATE

TIME (s)

0 50 100 150-1.5

-1

-0.5

0

THRUST REDUCTION

0 50 100 150

-4

-2

0

x 10-3

0 50 100 150

-2

-1

0

1 x 10

-3

0 50 100 150-2

-1

0

1

2 x 10

-4

TIME (s)

0 50 100 150-1.5

-1

-0.5

0

DRAG INCREASE

0 50 100 150

-4

-2

0

x 10-3

0 50 100 150

-2

-1

0

1x 10

-3

0 50 100 150-2

-1

0

1

2x 10

-4

TIME (s)

0 50 100 150-1.5

-1

-0.5

0

PITCH MEASUREMENT OFFSET

0 50 100 150

-4

-2

0

x 10 -3

0 50 100 150

-2

-1

0

1x 10

-3

0 50 100 150-2

-1

0

1

2x 10

-4

TIME (s) Figure 5: Fault signatures computed for the first stage ascend.

4.3 Estimation algorithm

Consider the linearized fault sensitivity model (3.19). In this model, )0;,,( ∆NtY b is the residual data

vector for the prediction based on the nominal model. In the absence of fault, the residuals should

be zero. This assumes the available prediction model ideally explains the data. In reality this is

never the case. For any real–life system the residuals would differ from identical zero because of the

sensor noise and modeling errors (scatter). All these factors are herein collected together in one

generalized ‘noise’ data vector. The following statistical model for the residual data vector Y is

assumed

eSfY += , (3.24)

Page 24: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

24

where S is a known sensitivity (fault signature) matrix and e is an unknown ‘noise’ vector.

By making certain statistical assumptions about the variables in the residual data model (3.24) it is

possible to arrive at the ‘optimal’ estimates of the fault parameters. Of course, in practice these

statistical assumptions should be taken with a grain of salt. One possible way to look at different

assumed statistical parameters is to consider them as tuning parameters of an estimation algorithm.

These parameters can be tweaked to achieve a reasonable practical performance of the algorithm.

With that in mind, we assume that the noise e in (3.23) is a normally distributed vector with a

covariance matrix V. We also assume that the unknown fault vector f being estimated is an

independent (of the noise e) random variable and has a normal distribution with covariance matrix

R. In that case an optimal statistical estimate of the fault vector f from the noisy data can be

obtained as a Maximal Likelihood (Maximum A posteriori Probability) Estimate in the form

[ ] YVSSVSRf TT 1111ˆ −−−− += (3.25) The estimate (3.23) is obtained for the residual vector Y using the sensitivity matrix S. Actually

computing the estimate (3.23) requires also having the covariance matrices V and R as well. The

estimate (3.25) is known as a GLS (Generalized Least Squares) estimate.

The noise covariance V can be empirically estimated from the residual scatter in the absence of

faults. The covariance V can be expressed via autocorrelations and cross-correlations of the residual

channel noise. In this work the data were generated in a simulation and contained no sensor noise.

The only modeling error present was the linearization error in the first-order approximation (3.9) of

the residual dependence on the fault vector. When implementing the estimation algorithm, the

covariance matrix V was selected as a scaled unity matrix. This corresponds to the white noise of the

same amplitude being present in all the observation channels.

The covariance matrix of the faults was chosen in the form

R = diag {(F1) 2, …, (F4) 2} , (3.27)

Page 25: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

25

where different faults are assumed to be probabilistically independent and Fj has a meaning of

expected fault intensity. The fault intensities can be set based on the knowledge of the fault nature

and were chosen as {F1, F2, F3, F4} = {20, 3.16, 10, 0.71}.

So far, the reasoning ignored the time dependence (time evolution) of the variables. In principle, the

estimate (3.23) can be obtained as a batch estimate from the entire set of data collected throughout

the flight. However, there might be a need to compute and update estimates through the flight so

that evolving faults can be estimated and observed on-line. To satisfy this need, we implemented

the GLS algorithm (3.25) for the fault parameter estimation in a receding horizon form. At each time

step, the receding horizon algorithms compute an estimate of the fault parameter vector using N

most recent data points. This estimate can be expressed in the form

[ ] ),,();0;,,()(ˆ 1111 ∆∆−=∆∆−+= −−−− NNtSSNNtYVSSVSRtf tTtt

Tt (3.27)

The result of computing the receding horizon estimate is shown in Figure 6 below. The four

subplots correspond to one fault each and show the seeded fault value (dashed line) and the fault

recovered by the receding horizon GLS algorithm from the data (solid line). In this example, the

data is sampled every 100 ms and the receding horizon algorithm uses a 250 point (25 s) history

buffer data for the residuals and fault sensitivities. The receding horizon estimate is updated every

10 s. This is visible as steps in the ‘staircase’ plots of the estimates. As one can see, the obtained

estimates are reasonable at each step and quite close to the seeded fault values. This is despite the

unmodeled nonlinearity of the fault dependence and a large condition number of the sensitivity

matrix. For example, at the last receding horizon computation step the condition number is 3.8⋅104.

Page 26: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

26

0 50 100 150 0

10

20

30 GIMBAL SLUGGISHNESS, PERCENT

0 50 100 150 0

0.5

1

1.5 THRUST REDUCTION, PERCENT

0 50 100 150 0

1

2

3

4

5 DRAG INCREASE, PERCENT

0 50 100 150 0

0.2

0.4

0.6

PITCH MEASUREMENT OFFSET, PERCENT

Seeded Fault

Receding Horizon Estimate

Figure 6: Receding horizon GLS estimate of the faults.

A few extensions of the implemented estimation algorithm are possible. The receding horizon

estimation can be posed as a constrained optimization problem that incorporates a priori knowledge

about possible fault magnitudes as the optimization constraints. It could be also possible to

incorporate the knowledge of the fault evolution. In particular, one piece of such knowledge is that

some of the fault parameters can only evolve one way – the mechanical damage is accumulating

and no ‘healing’ is happening. A monotonic regression algorithm for fault estimate trending based

on constraint optimization is discussed in [6].

5. Conclusions

The paper has described a case study for centralized parametric diagnostics of flight critical systems

of a launch vehicle. The demonstrated architecture partitions model-based computation of the

Page 27: Parametric Diagnostics of Flight Control and Propulsion for …gorin/papers/AIAA_JGCD_IVHM_04.pdf · 2004-06-21 · technology was subsequently used in Honeywell’s Primus Epic CMC

27

residuals and faulty signatures into vehicle subsystem processors. The residuals are integrated for

the estimation in the central IVHM processor, which does not require access to the detailed

modeling knowledge. A receding horizon algorithm for GLS fault estimation was described. This

algorithm was demonstrated to provide accurate fault estimation in launch vehicle ascent

simulation.

6. Acknowledgement The authors wish to acknowledge useful discussions with and valuable application description

contribution of Carlos Garcia-Galan.

7. References

1. A. Alessandri, M. Baglietto, and G. Battistelli,, “Receding-horizon estimation for discrete-

time linear systems,” IEEE Tran. on Automatic Control, Vol. 48, No. 3, 2003, pp. 473--478

2. R.N. Clark, P.M. Frank, and R.J. Patton. Advanced is Fault Diagnosis for Dynamic Systems.

Springer Verlag, 2000

3. P.M. Frank ``Fault diagnostics in dynamic systems using analytical and knowledge-based

redundancy - A survey and some new results,'' Automatica, Vol.26, No.3, 1990, pp. 459--474.

4. S. Ganguli, S. Deo, and D. Gorinevsky, “Parametric fault modeling and diagnostics of a

turbofan engine,” IEEE Conf. on Control Applications, September 2004

5. J. Gertler ``Survey of model-based fault isolation and detection in complex plants,'' IEEE

Contr. Syst. Magazine, 1988.

6. D. Gorinevsky, “Monotonic regression filters for trending deterioration faults,” American

Control Conference, Boston, MA, June 2004

7. R. Isermann and P. Balle “Trends in the application of model based fault detection and

diagnosis of technical processes,” 1996 IFAC World Congress, San Francisco, CA, July 1996

8. K. R. Muske, J. B. Rawlings, and J. H. Lee, “Receding horizon recursive state estimation,”

American Control Conference, June 1993, pp. 900—904.

9. G. Ramohalli. “The Honeywell on-board diagnostic and maintenance system for the Boeing

777,” 11th IEEE/AIAA Digital Avionics Systems Conf., pp. 485--490, October 1992