You are on page 1of 33

Accepted Manuscript

Trustable UAV for Higher Level Control Architectures

Chimpalthradi R. Ashokkumar, George WP York, Scott Gruber

PII: S1270-9638(17)30827-1
DOI: http://dx.doi.org/10.1016/j.ast.2017.05.013
Reference: AESCTE 4026

To appear in: Aerospace Science and Technology

Received date: 9 February 2016


Revised date: 12 October 2016
Accepted date: 4 May 2017

Please cite this article in press as: C.R. Ashokkumar et al., Trustable UAV for Higher Level Control Architectures, Aerosp. Sci. Technol.
(2017), http://dx.doi.org/10.1016/j.ast.2017.05.013

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing
this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is
published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
Trustable UAV for Higher Level Control Architectures
Chimpalthradi R. Ashokkumar1, George WP York2 and Scott Gruber3

Abstract – Future unmanned air vehicles (UAVs) are expected to operate under the supervision

of higher level control architectures that give instructions to engage and direct the UAV to

perform a certain task. These instructions allow the UAV to make decisions along any point of

its trajectory and then modify its flight path by using a sequence of reconfigurable controllers at

the decision points. Assume that the UAV is flying with a transient and with a steady state

contributing to a flight control mode (FCM) such as an altitude hold, an ascent and a descent

mode when only longitudinal aircraft dynamics is considered. At the time of a decision, if the

UAV bifurcates from its original (or parent) FCM in an effort to acquire a new (or a child) FCM

and comply with a higher level instruction, then the UAV is said to be trustable. Mathematically,

at the time of bifurcation where controller reconfiguration takes place, trustable UAV with the

child trajectory must originate from a region where the stability regions of the parent and child

trajectories intersect. In this paper, a procedure to reconfigure such FCMs and their controllers

that develop the trustable UAV are presented. A three degree of freedom UAV is considered to

illustrate the trustable UAV.

Keywords: Flight control modes, UAVs, transients, steady states, higher level instructions,

switching, stability.

1
NRC Research Associate, Dept. of Electrical and Computer Engineering, USAF Academy, Colorado, USA.
2
Director and Associate Professor, USAF Academy Center for Unmanned Aircraft Systems Research, Dept. of
Electrical and Computer Engineering, USAF Academy, Colorado, USA.
3
Program Manager, USAF Academy Center for Unmanned Aircraft Systems Research, Dept. of Electrical and
Computer Engineering, USAF Academy, Colorado, USA.

1
1. Introduction

Future unmanned air vehicles (UAVs), both aerial and combat air vehicles, have a major role

to mitigate threats involving the life of ground personnel and pilots who are assigned to perform

complex mission operations. In these operations, although human decisions to remotely operate

an UAV are far superior, the UAVs are expected to be operative under the supervision of a

higher level control architecture to carry out a task. The instructions are such that they prompt

the UAV to make a decision and change the course of its flight path. At the decision points, if the

UAV is able to reconfigure its flight path with stability, then the UAV is said to be trustable.

Otherwise, the UAV at that decision point is not trustable. In this paper, a mathematical

exposition using the linearized model based controllers of the nonlinear UAV that develops a

trustable UAV is presented. The trusted UAV is assumed autonomous. The technical factors

governing the human operated UAV will be considered at a later stage to define the trust.

Nevertheless, pilot errors [1] in manned or unmanned aircraft leading to catastrophes require

special attention to define trust. The aircraft simulation in these cases must be carried out like the

way the pilot with an inner loop controller switching criterion operates the aircraft.

In unmanned aircraft, there is a requirement to perform nonlinear aircraft simulations using

switched linear model based controllers that guarantee trust, especially when higher level control

architecture’s instructions [2,3] decide to direct an extremely uncertain lower level UAV [4,5] to

perform a certain task. Typical uncertainties influencing the trustable UAV are higher level

decision points at which the UAVs are instructed, denied airspace, friendly resources, enemy

states, etc. Such uncertainties constantly drive the aircraft to adopt an unplanned flight path. One

option available to validate a trustable UAV in these circumstances is to develop and automate a

capability to switch flight control modes (FCMs) such as an ascent, descent, and altitude hold,

2
etc., compatible to higher level instructions so that the flight bifurcates from its original (or

parent) FCM to a new (or a child) FCMs as shown in Figure 1.

Fig. 1. A trustable UAV with parent and child FCMs

The parent FCM shown in solid line (ABCD) has a transient and a steady state. The child FCM

shown in dotted lines can originate from any part of the parent’s transient or the steady state

depending upon the time at which a higher level instruction is directed. The solid and dotted line

combination depict the trustable UAV, which takes such instructions at nodes and makes

decisions to bifurcate instantaneously from the parent FCM with a FCM reconfiguration for the

child. In this paper, a procedure to develop these trustable UAVs is presented. In particular,

controller reconfiguration to generate parent and child FCMs in extended stability regions is

presented. That is, the trustable UAV is developed by using controller switching as in gain

scheduling algorithm; however, these controllers are developed to originate from a particular

trim point of the flight envelope. Following the control and simulation procedure, the trustable

UAV possesses the inner loop controller switching capability and guarantee stability of the

reconfigurable FCMs that the parent and child generate. For a given command or pilot input at

the outer loop, the rules to determine the sequence of controllers to be switched as in gain

scheduling algorithms [6-11] consistent with flight envelope points are beyond the scope of the

present work. Hence, in this paper, trustable UAVs are investigated with respect to the initial

3
conditions and various controllers, all originating from a single linear time invariant model (or

the trim point).

It is possible to develop a trustable UAV whose FCMs are derived by considering various

flight conditions on the flight envelope. Here, the linearized model with respect to its equilibrium

or trim point varies for each flight condition. Controllers are determined for each linear model

and they are scheduled as per the norms of the gain scheduling algorithm. Note that controller

switching needs to take place at a point where the stability regions (regions of attraction) of the

respective controllers intersect [12]. In current gain scheduling algorithms, the stability regions

are so intact and the switching criteria are therefore taking place within a region where stability

regions intersect. Alternative approach to avoid switching rules is simultaneous stabilization

[13,14]. Here a single controller attempts to stabilize as many linear models as possible;

however, such controllers are difficult to determine for more than two linear models [15,16].

While there is a need to understand the selection of flight conditions in accordance with a flight

trajectory, in this case FCMs, it is also possible to override controller variations with altitude

(one of the coordinates of the flight envelope) where their linear models simultaneously vary.

For instance, a sequence of controllers pertaining to a single linear model can generate ascent

and descent modes as against the gain scheduling algorithm where controllers and their linear

models are assumed to vary with altitude. For UAVs, these procedures are open for

implementation. Along these lines, presently, sophisticated six degree of freedom nonlinear

aircraft and its equations for navigation is not considered. This paper assumes a single

longitudinal linear model to generate parent and child FCMs for a trustable UAV. The child

FCMs are reconfigured by using controllers resulting from various closed loop pole locations.

Presently, the linear model corresponding to steady level flight of a longitudinal micro air

4
vehicle model [17] is considered. Generally, FCM variations are possible with thrust and

elevator as control inputs. Hence the linear model becomes a two input system.

FCMs of the nonlinear aircraft and eigenvalue and eigenvector (eigenstructure) options for

the linear model based controller design are strongly connected; however, eigenvalue selection

or the choice of eigenvector elements for a particular ascent or a descent mode is difficult to

establish. In fact, in aircraft applications, the choice of eigenvector elements for eigenmode

decoupling [18], minimum norm gains for control input constraints [19], etc., does not apply

when the controller is implemented for the nonlinear aircraft. For each set of desired eigenvalue

locations, eigenvector options are indeed infinite. Accordingly, FCM options become infinite. A

procedure to determine such FCMs in state feedback format [20], in output feedback format [21],

and in observer based feedback format [22] is presented. Potential steady state trajectory

transcriptions for stability of reconfigurable controllers in an integrated FCM to depict a typical

reconfigurable aircraft control are discussed. In this paper, in order to develop a trustable UAV,

both transient trajectory transcriptions (TTTs, for higher level instructions at transients) as well

as steady state trajectory transcriptions (SSTTs, for higher level instructions at steady states)

using controllers resulting from different closed loop eigenvalue locations are considered.

Eigenvalue variations in a way depict variations in eigenvectors. Each of these choices will have

a unique steady state value called auxiliary equilibrium point to define the FCM and several

transients depending upon the initial condition one picks up from the stability region. If a parent

and its associated child FCM is required to be configured, initial condition (or transient) for the

parent FCM becomes instrumental to have acceptable higher level decision points where the

child FCMs emerge. Such conclusions are also derivable by using the controllers resulting from

robust pole assignment [23], linear quadratic regulator [24], etc. The basic objective behind a

5
trustable UAV is that the bifurcation from parent to child FCM is stable and it is a property

associated with the variations in eigenstructure of the linear model. Eigenstructure variations can

be pursued through linear functional controllers (LFCs) [20-22] (the eigenvalues are retained at

fixed locations but their eigenvectors are varied through a multiple input state variable feedback

controllers). In this paper, the basic requirement on eigenvector variations for FCM options is

obtained through eigenvalue variations.

The rest of the paper is organized as follows. The problem description is presented in Section 2.

Trajectory transcription criteria for a trustable UAV is presented in Section 3. In Section 4, the

autonomous attributes of the trustable UAV are discussed. Simulations and conclusions are

provided in Section 5 and 6, respectively.

2. Problem description

When a higher level instruction to engage an UAV and perform a certain task is made, it is

important that the UAV need to cooperate with the instruction and make a decision to develop a

child trajectory. Several scenarios exist to define the present and future state of the UAV at the

time the decision is made. For instance, prior to decision, the UAV may be accelerating to a new

altitude in an ascent mode and after the decision it may be required to pass through a waypoint

(say a latitude, longitude and altitude point). A question that would naturally arise is the

following. Can the UAV bifurcate from its present path to its future path? As in standard aircraft

control problem, when the inner loop is scheduled with linear model based controllers, this paper

considers the FCMs of the longitudinal aircraft dynamics and proposes parent and child FCMs

for bifurcation at the decision points. As discussed in the previous section, the controllers work

like the gain scheduling algorithm but the trim points are not. That is, several pole placing

controllers designed for one linear model are scheduled to derive the child FCMs.

6
At the inner loop, LFCs presented in [20-22] are particularly attractive to understand FCM

transcriptions as a function of condition numbers for the closed loop modal matrices. Here, the

respective eigenvalues in all the FCMs are fixed in the open left half plane of the complex plane.

However, the eigenvectors are varied infinitely to have variations in condition numbers. In this

paper, the eigenvalue locations are varied and the procedure to design these LFCs for variations

in condition numbers is retained. Hence the FCM options are multiplied for every set of desired

eigenvalue locations. Further, the FCM transcriptions and hierarchical controllers that assign

distinct set of eigenvalue locations which transcript FCMs and configure parent and child FCMs

are unknown and it is the study objective of this paper. Such an assignment can also take the

controllers resulting from robust pole assignment [23] and linear quadratic regulators [24]. In

order to define a FCMs and their transcriptions for a trustable UAV, the nonlinear aircraft

confining to the pitch plane [17] is taken in the form,

x a( x)  b( x)u(t ) . (1)

Assume the control input u(t) is not a scalar. The n-component state vector x(t) requires the

compatible vector valued functions a(x) and b(x) to be smooth and differentiable with respect to

x(t). The time (t) derivative of the state variables is denoted by a dot over the state vector x(t).

Let the kinematics of the aircraft in NED (North-East-Down) coordinate frame as,

X i va (t ) cos(T (t )  D (t )) (2)

Zi va (t ) sin(T (t )  D (t )) (3)

where va (t ) is the velocity of the aircraft in m/s, T (t ) is the pitch angle in radians and D (t ) is the

angle of attack in radians. The flight paths are determined by X i and Z i , which are the

coordinates of the inertial positions such as North-Down or East-Down positions. Other state

(trajectory) and control variables that govern the dynamics of the aircraft in pitch plane are pitch

7
rate ( T Q (t), rad/sec), elevator deflection ( G e (t ) , rad) and thrust coefficient CT (t ) . The state

vector is identified as follows.

xc(t ) >va D T Q@ (4)

The control variables are identified as,

u1 (t ) CT (t ) and u2 (t ) G e (t ) . (5)

The three degree of freedom micro aerial vehicle model and its aerodynamic data is given in

[17]. An equilibrium point (in Eq. (1), points where x 0 ) corresponding to a steady level flight

is computed. The nonlinear UAV from [17] is given by,

S S S
va (t ) q f LD (CL0  CLD D )  g sin (T  D )  q CT cos D  q CDG G e . (6a)
m m m e

1 S Sc S g
D (t ) [Q  q CL0  q 2 CLQ Q  q CLD D  cos(T  D )]
k va m 2 va m v a m v a
(6b)
S S
q CT sinD  q CL G e
k va m k va m Ge

Sc c Sc
T(t ) q [Cm0  CmD D  CmQ T]  q CmG G e (6c)
I yy 2v a I yy e

where,

f LD (M ) a1M 4  a2M 3  a3M 2  a4M  a5 , M CL0  CLD D


(6d)
a1 0.1723; a 2 0.3161; a3 0.2397; a4 0.0624; a5 0.0194

Sc
k 1 q CL , (6e)
2va2m D

q 1
2
Uva2 , (6f)

Here dot over a variable refers the time derivative of the variable. U is the air density kg/m3, S is

the surface area of the wing in m2, m is the mass of the aircraft in kg, c is the mean aerodynamic

chord in meters, Iyy is the pitch moment of inertia in kg-m2 and other parameters are the

aerodynamic stability and control derivatives. The data for these coefficients are given in [17].

8
To compute the trim condition, special techniques are required to solve a set of nonlinear

algebraic equation with six unknowns and four equations (for six degree of freedom aircraft, see

[25]).

Fig. 2. Potential parent red) and child (blue) FCMs of an UAV.

In order to explain parent and child FCMs, consider the transcriptions (FCMs in blue

transcripting FCMs in red) as in Figure 2. Each slope is a representation of the inner loop

controller. The transient is a representation of the initial condition belonging to its stability

region. Further, the transients are confined to admissible state variable bounds which in turn

determine bounds on the control inputs. The points where the color codes change and trajectories

in blue and red intersect are the decision points where controller reconfiguration takes place to

oblige the higher level instructions. Note that the decision points are not observed on transient

9
part of the trajectories. Suppose the higher level instructions come in at any point of the

trajectory. When a particular trajectory in red (referred as parent trajectory or FCM) with which

the trajectories in blue (referred as child trajectories or FCMs) transcript at many points of the

parent trajectory, the UAV becomes trustable. Depending upon the controller for parent

trajectory, the child FCMs are defined. For instance, an extremely shallow parent will not have

transient effects and hence the child FCMs are derivable at almost every point of the parent

trajectory. The parent trajectory in this case is a property associated with the initial condition

search in the stability region of its controller. Given this controller, a procedure to generate the

parent trajectory using a linear time invariant model based controller is well known in the

literature. Usually, the aircraft in Eq. (1) is linearized with respect to an equilibrium point. The

linear model is completely controllable and is given by,

'x A 'x  B 'u(t ) 'x x (t )  x e 'u u( t )  ue (7)

In the present case, a LFC G defines a control law of the form,

'u(t ) G 'x(t ) (8)

For each of these controllers ( G i ), the trajectories x(t ) in Figure 2 are generated from the

following differential equations (for brevity, the superscript i is ignored),

x a( x)  b( x)[ue  G ( x  xe )] , x(t 0) z 0 . (9)

Note that the auxiliary equilibrium points xai ,e at FCM transcriptions are the points that the

numerical integration scheme employed in navigation and control algorithm satisfy,

a( xai ,e )  b( xa(i,)e )[ue  G i ( xai ,e  xe )] 0 , i 1, 2  (10)

Also, when the auxiliary equilibrium point is too close to the equilibrium point, transcription is

favored by the fact that it will be in a region where stability regions of the LFCs intersect, a

condition investigated in detail for a second order nonlinear system [12]. A procedure for such

10
transcriptions is presented in sequel and it holds for state feedback based LFCs when the

controllers are originating from the model in Eq. (7). These LFCs for state feedback case is

presented in [20]. In addition, the state feedback controllers resulting from robust pole

assignment and linear quadratic regulators are also considered which primarily distinguishes the

slopes of the FCMs from other techniques. Prior to presenting transient trajectory transcriptions

(TTTs) and steady state trajectory transcriptions (SSTTs) by using a combination of stability

regions pertaining to the LFCs, the stability region is defined as follows:

Definition 1 (Stability Region):

SGj ^x
0  R 4 : If x(0) x0 then limtof x(t ) `
xaj,,e , j represents the linear functional controller G j .

Note that the classical stability region definition takes the equilibrium ( xe ) and auxiliary

equilibrium ( xaj,e ) points as xe xaj,e 0 . In aircraft, these non-zero points are crucial to explain

trustable UAV, where the controllers are switched to define child FCMs from a parent FCM.

Consider now one of these trustable UAVs and denote its controller for a parent FCM by G p and

its controllers for child FCMs by G d , d 1,2 . The TTT is defined as follows:

Definition 2 (Transient trajectory transcriptions TTTs):

Consider a stability region S Gp for a parent controller G p . Denote a trajectory in the stability

region S Gp by x p (t ) . Similarly, consider a stability region S Gd for a child controller G d . Let the

decision point on the parent trajectory be t d , which is the time at which the UAV with a child

trajectory begins to bifurcate from the parent trajectory. Then the UAV with a TTT is contained

in the child stability region S Gd with the following criterion.

S Gd ^x d
p  R 4 : If x p (t td ) x dp then lim t t t d of x(t ) `
x ad,e , G represents a child controller G d .

Similarly, the SSTTs for decision points on auxiliary equilibrium points are defined.

11
Definition 3 (Steady state trajectory transcriptions SSTTs):

Consider a stability region S Gp for a parent controller G p . Denote the auxiliary equilibrium point

of the trajectory x p (t ) in the stability region S Gp by xap,e . Similarly, consider a stability region S Gd

for a child controller G d . Let the decision point on the parent trajectory be t d , which is the time

at which the UAV with a child trajectory begins to bifurcate from the parent trajectory. Then the

UAV with a SSTT is contained in the child stability region S Gd with the following criterion.

S Gd ^x p
a ,e  R 4 : If x p (t td ) x ap,e then limt tt d of x(t ) `
x ad,e , G represents a child controller G d .

One of the implications of the definitions provided for TTTs and SSTTs is that the stability

regions S Gp and S Gd must intersect so that x dp and x(t d ) (parent and child trajectories at the

decision point) reside at the intersected region. These definitions are used to define the FCM and

the trusted UAV as follows:



Definition 4 (Flight Control Mode): In the stability region ܵீ (or ܵீௗ ), consider an initial

condition ‫ݔ‬଴ (or ‫ݔ‬ሺ‫ݐ‬ௗ ሻ) such that the trajectories contained in their respective stability regions
௣ ௗ
approach the steady state ‫ݔ‬௔ǡ௘ (or ‫ݔ‬௔ǡ௘ ). The FCM with an ascent (+) or descent (െ) will be

௣ ௣ ௗ ௗ
േ–ƒሺߠ௔ǡ௘ െ ߙ௔ǡ௘ )) (or േ–ƒሺߠ௔ǡ௘ െ ߙ௔ǡ௘ )).

Definition 5 (Trustable UAV) : If the UAV trajectory ‫ݔ‬ሺ‫ݐ‬ሻ comprising of parent and child

trajectories is contained within the union of the stability regions ܵீ௣೛ ‫ீܵ ׫‬ௗ೏ such that the point

‫ݔ‬ሺ‫ ݐ‬ൌ ‫ݐ‬ௗ ሻ ൌ ‫ݔ‬௣ௗ resides at the intersected region ܵீ௣೛ ‫ீܵ ת‬ௗ೏ , then the UAV is said to be trustable.

Note that the parent and child controllers ‫ ܩ‬௣ and ‫ ܩ‬ௗ , respectively, are reconfigurable at ‫ݔ‬௣ௗ .

In order to know if the SSTT and TTT criteria for a trustable UAV are satisfied, it becomes

necessary to interpret Lyapunov stability criterion for nonlinear autonomous system using

eigenvalues of the Jacobian matrix. This technique also accommodates the LFCs G d and G p that

12
are designed for eigenvector variations which are required to have options in auxiliary

equilibrium points xad,e that in turn decide ascent and descent modes of the child FCMs. Some of

these principles are discussed in sequel.

3. Transcription criterion in a trustable UAV

The nonlinear aircraft in Eq. (9) is indeed the nonlinear autonomous system of the form,

x f (x ) , x(t0 ) x0 z 0 (11)

and one of its non-zero equilibrium points xe such that

f ( xe ) 0 (12)

As in [26,27], choose the Krasovskii’s Lyapunov function V ( x) ! 0 such that

V ( x) 1
2
f cP f , P ! 0

where prime denotes the transpose. The time derivative of the Lyapunov function is,

wf ( x )
V ( x) f c ^Ac( x) P  PA( x)` f , A(x ) is Jacobian . (13)
wx

Clearly, for stability, V ( x)  0 implies that A(x ) is asymptotically stable along the trajectory x(t ) .

In [26,27] it has been shown that Eq. (13) has a Lyapunov equation connection [28].

Ac( x) P  PA( x) Q , P ! 0 and Q t 0 . (14)

Theorem 1: The equilibrium point x e is asymptotically stable if A(x ) is in Ф, where Ф is the set

of all matrices whose eigenvalues have negative real parts.

In order to connect condition numbers and transcriptions, Theorem 1 is used. Consider the LFCs

G (i ) and define the closed loop system matrices A0(i ) such that,

A0(i ) A  BG (i ) , i 1, 2, (15)

Theorem 1 immediately defines unstructured error matrices E (i ) ( x) such that,

A(i ) ( x) A0(i )  E (i ) ( x) , i 1, 2, (16)

13
The eigenvalue perturbations of A(i ) ( x ) from the nominal eigenvalues of A0(i ) are connected as

follows.

Theorem 2: Let the nominal matrix A0(i ) is diagonalizable such that A0(i ) V (i ) /V (i )1 where V (i )1

is the inverse of V (i ) and / diag (O1 On ) . If an error matrix that perturbs the nominal matrix

A0(i ) is E (i ) ( x) and if P (ip ) is the eigenvalue of A(i ) ( x) A0(i )  E (i ) ( x) , then

min P (pi )  Oq d N (V (i ) ) V max ( E (i ) ( x)) . (17)


1d q dn

Theorem 2 is well known when x(t ) and i are fixed [29,30]. Here N (.) refers the condition

number of the matrix (.). Similarly, V max (.) refers to the maximum singular value of the matrix

(.). Note that in Eq. (17) the property of the LFC for eigenvector variations when eigenvalues are

fixed or eigenvalue variations is employed. In order to present the transcription criterion, few

inferences of Theorem 2 are presented as follows.

A. Smaller the condition number larger the stability region.

Consider the left hand side of Eq. (17). The perturbed eigenvalue P q(i ) will be within a circle of

radius ‘R’ with center at Oq if [30],

R
V max ( E (i ) ( x )) d (18)
N (V (i ) )

Clearly, smaller the condition number larger the error size E (i ) ( x) . That is, larger the trajectory

dispersions (that is, x(t) – dispersions) from the equilibrium point xe at which E (i ) ( xe ) 0 . Hence

stability region is larger when condition number for the modal matrix of the nominal matrix A0(i )

is smaller. Note that the circles with radius ‘R’ and with centers at the eigenvalues of the

nominal matrix are contained in the open left half plane of the complex plane. The transcription

criterion in a trustable UAV is immediately inferred as follows.

B. Transcription criterion.

14
It is observed that all the controllers for parent and child FCMs originate from the linear model

in Eq. (7) whose equilibrium point is given by ( xe , u e ). The Jacobian matrix A(i ) ( x) in Eq.(16) at

this equilibrium point is indeed the nominal matrix A0(i ) for which the error matrix E (i ) ( xe ) is a

zero matrix. Hence when the Jacobian matrix at all x(t ) is asymptotically stable, the error matrix

E (i ) ( x) measures the relative position of x(t ) with respect to the equilibrium state xe . If one takes

the 2-norm of the error matrix E (i ) ( x) denoted by V max ( E (i ) ( x)) and take this norm for the parent

and child trajectories at the decision point t d , transcription criterion is inferred as follows. For

the child trajectory x(t d ) to transcript with parent trajectory x dp , a qualitative requirement is that,

V max ( E ( p) ( x dp ))  V max ( E (d ) ( x(t d )) (19)

Eq. (19) suggests that at the decision point, the stability region of the child controller most likely

contains the stability region of the parent controller enabling the potential TTT. The same

principle applies to SSTT where for the child FCM to most likely transcript the parent FCM, the

condition is established as follows.

V max ( E ( p) ( xap,e ))  V max ( E (d ) ( x(t d )) (20)

These qualitative rules of transcription criterion assist to accommodate the decision points of the

higher level control architecture while an UAV is directed to perform a certain task. Eqs (19) and

(20) further suggest (qualitatively) that when a decision point is active, the distance of the child

trajectory point from the equilibrium point is more compared to the distance of the parent

trajectory point from the same equilibrium point (as in Figure 3). These qualitative statements

are used for a transcription to occur.

15
Fig. 3a. Child controller selection criterion.

Fig. 3b. Stable child trajectories for higher level decisions.

C. LFC effects on transcriptions.

LFCs fundamentally adopt eigenstructure variations to modify transients and steady states in

their respective stability regions. Such variations may be established by designing a controller

whose closed loop eigenvalues are either fixed or varied in the open left half plane of the

complex plane. The freedom available to design these controllers via eigenstructure options

offers flexibility to access any decision point of the higher level control architecture; however,

identifying a controller for child FCM that satisfies the transcription criterion (Eq. (19) or Eq.

(20)) becomes difficult. In this framework, it can be claimed that all decision points of the higher

level control architecture are accessible through appropriate child controllers. Suppose the child

controller is fixed. Again it is observed that the initial condition options suggest different

16
decision points at transients. Hence it becomes important to highlight the significance of the

point x(t d ) on the child trajectory.

Theoretically, the UAV at the decision point t d begins to acquire a child trajectory with an initial

condition x dp on the parent trajectory. While it is known that x dp is in the stability region S Gp for

stable switching of the controller G d from G p . In order to check if x dp is in the stability region

S Gd of the controller G d , the following hypothesis test is applied. Let x0 be an initial condition

that generates trajectories x p (t ) in S Gp and x(t ) in S Gd . Clearly if x(t d ) satisfies the trajectory

transcription criterion in Eq. (19), it is likely that x dp is in S Gd . The hypothesis test is a procedure

to choose potential controllers for child FCMs. The entire procedure selects the initial condition

x dp for the child FCM generated by a controller option G d . It is anticipated that the Lyapunov

stability criterion in the form of an asymptotically stable Jacobian at Eqs. (13) and (14) will be

satisfied for an f (x) satisfying,

x f (x) , x(t d ) x dp , where (21)

f ( x) a( x)  b( x) [ue  G d ( x  xe )]

In Figure 4, a detailed description of a trustable UAV is presented.

4. Autonomous attributes of a trustable UAV

Before illustrating the autonomous attributes of a trustable UAV, the current practices in UAV

operations by a remote operator are reviewed. Figure 5 depicts the typical control architecture

with a remote operator at the outer loop. The main feature of the architecture is that at the inner

loop a fixed order linear control structure with constant gains is adopted. If it is a PID controller,

the gain tuning is performed usually by trial and error method. If a FCM such as an ascent is

desired, the remote operator intuitively gives a command input for which the UAV ascends with

17
Fig. 4. Trustable UAV complying the higher level instructions.

a constant thrust. Note that from an affixed order inner loop controller, these are the limited

maneuver capabilities from an UAV without the thrust management. However, human decisions

significantly contribute to the feasible mission objectives of UAV. In contrast, the control

architecture of an autonomous trustable UAV for higher level decisions is depicted in Figure 6.

These architectures are programmable if the number of parent and child controllers is known.

The decision points may be appropriately directed; however, the difficulty prevails for decision

points at transients. In this regard, LFC may be designed such that,

V max ( E p ( x p (t ))  V max ( E d ( x(t )) (22)

Here, the higher level decisions can be made at any part of the parent trajectory for all of which

the transcription criterion will be met to generate child FCMs. In the next section, the

18
autonomous features of a trustable UAV presented in Figure 5 are simulated by using the parent

and child FCMs.

Fig. 5. FCM in a human operated UAV.

Fig. 6. Autonomous parent and child FCMs in a trustable UAV.

19
5. Simulations

The three degree of freedom aircraft model of Langelaan [17] is considered to simulate a

trustable UAV. The aircraft is actually a four-state and three-input system. In this paper, the

utility of the flap as an input is not investigated. The state and other control variables defined in

Eq. (4) respectively are recalled as follows.

ªva (t )º ªForward velocity, m/sº


« D (t ) » « Angle of attack, rad »
x(t ) « » « ».
« T (t ) » « Pitch angle, rad »
« » « »
¬ Q(t ) ¼ ¬ Pitch rate, rad/s ¼

ªCT (t )º ª Thrust coefficien t º


u (t ) « G (t ) » «Elevator deflection , rad » .
¬ e ¼ ¬ ¼

The kinematics of the aircraft in pitch plane are given in Eqs. (2) and (3). The position of the

aircraft is identified by ( X i (t ), Z i (t ) ) as in Figure 2, where X i (t ) is the latitude longitude position

in meters and Z i (t ) is the altitude in meters.

One of the simplest equilibrium points considered to develop the trustable UAV is computed as,

xec >21.021 0 0 0@ and uec >0.0163 0@

The aircraft at this equilibrium point is in cruise (altitude hold) mode with D (t ) T (t ) 0 . For a

cruise mode with non-zero angle of attack, generally an optimization is required to compute the

associated equilibrium point and it is illustrated in [25]. For Eq. (23), the controllable linear

model is computed as (all data are approximated to the fourth decimal place),

20
ª 0 6.7890  9.8100 0 º
« 0.0452  7.2167 0 1.0280 »»
A «
« 0 0 0 1.0000»
« »
¬ 0  34.3897 0  2.7163¼

ª27.1758 0 º
«0 0.4806 »»
B «
«0 0 »
« »
¬0 54.8718¼

Following the procedure presented in [20], the state feedback LFCs G (i ) are computed for

eigenvector variations of the closed loop system matrices Ac(i ) A  BG (i ) . For instance, when

eigenvalues of Ac(i ) are fixed at,

5 r 5.5 j
 2 r 2.5 j

two LFCs out of infinite options are computed as,

ª0.1552 0.5727  0.7073  0.1845º


G1 «1.1461  3.2203  2.2719 0.0254»
¬ ¼

ª 0.1108 0.8881  0.2090  0.0405 º


G2 «  0.8533 1.1311 0.6653
¬ 0.1191»¼

Other eigenvector variations are pursued by changing the eigenvalues. For,

4 r 7.5 j
,
 2 r 2.5 j

one of the controllers assigning the above closed loop poles is,

ª 0.0054  0.1911  0.2758 0.1415º


G3 « 0.9719  1.5841  1.2641 0.0489»
¬ ¼

For,

3 r 7.5 j
,
 2 r 2.0 j

one of the controllers assigning the above closed loop poles is,

21
ª 1.3607 7.0370 2.2027 1.1524º
G4 « 0.7126 2.5202 1.2994 0.6348» .
¬ ¼

Table 1 Auxiliary equilibrium points of S Gj

xaj,e

ª v aj,e º
« j »
Controller «D a ,e » dZ i
«T j »  tan(T aj,e  D aj,e )
« aj,e » dX i
«¬Qa ,e »¼ FCM
G 1
ª 20.9683 º
« 0.0014»
« »
« 0.0250»
« »
¬ 0.0000¼ 0.0237 Ascent Mode
G2 ª 21.0259 º
« 0.0017 »
« »
« 0.0108 »
« »
¬ 0.0000 ¼ 0.0125 Descent Mode
G 3
ª 21.0362 º
« 0.0018»
« »
« 0.0103»
« »
¬ 0.0000 ¼ 0.0085 Ascent Mode
G 4
ª 21.0451 º
« 0.0019»
« »
« 0.0177 »
« »
¬ 0.0000 ¼ 0.0196 Descent Mode

Given these stabilizing controllers, the auxiliary equilibrium points xaj,e of the stability regions

S Gj are presented next. For any initial condition x0 in S Gj , the trajectory satisfying Lyapunov

stability criterion derived in terms of stable eigenvalues of the Jacobian matrix in Eq. (13) will

determine these auxiliary equilibrium points. Such points depict the slope of the FCMs in ascent

and descent modes. These details of the controllers G j are presented in Table 1. In order to

accommodate the decision points of the higher level control architecture, which direct the UAV

to bifurcate from its parent trajectory, the transcription criterion in Eqs. (19) or (20) is adopted. It

22
basically suggests controller switching within a region where their stability regions intersect, a

rule exhaustively illustrated in [12] for a simple second order nonlinear system. It is also

important to observe that there are state trajectory transcriptions as well as control input

trajectory transcriptions at the decision points. With respect to a transient (initial condition)

option pertaining to a stability region S Gj , the state trajectory transcription is addressed first for a

fixed initial condition x0 , which is taken as,

ª21.0210 º
« 0.0002»
x0 « ».
«0.0001 »
« »
¬0.0001 ¼

To identify parent and child FCMs, the norm of the error matrix in Eq. (18) is plotted for three

controllers taken at a time. Figure 7 is utilized to depict transcription criterion in Eqs. (19) and

(20). Controllers G1 (indicated in black color), G 3 (blue) and G 4 (red) are considered. Controller

‫ ܩ‬ଶ is not included as it will be used to illustrate the effect of controller combinations in

transcriptions subsequently. It is possible to select as many controllers as one can and illustrate

transcription criteria. For instance, in Figure 7, controller G 3 offers a trajectory which is likely to

contain in the most parts of the stability regions S G1 and S G4 of the controllers G1 and G 4 ,

respectively. Hence the trajectory resulting from controller G 3 becomes a parent trajectory. The

resulting FCM is referred as parent FCM. The fact that it is contained in S G1 and S G4 further

enables the controllers G1 and G 4 to transcript with controller G 3 . Upon satisfying the Lyapunov

stability criterion (Eq. (13) and (14)), the resulting trajectories and FCMs from these

transcriptions are referred as child trajectories/FCMs. Note that in steady state and at other parts

23
of the transients, other conclusions can also be drawn from Figure 7. That is, unlike the first part

Fig. 7. Transcription criterion depicting potential parent and child FCMs.

of the transients, at steady states, G1 transcripts with G 4 (as against G 4 transcripting with G1 at

the first part of the transients). This observation is useful if the child FCM becomes a parent

FCM at some decision points. To illustrate the main properties of transcription, consider a

decision point t d 3 seconds. From Figure 7,

V max ( E1 ( x(t d ))) 0.9976

V max ( E 3 ( x(t d ))) 0.1606 (23)

V max ( E 4 ( x(t d ))) 0.6338

24
Clearly, the point x(td ) on the trajectory resulting from the controller G 3 (denoted by x dp ) is

likely to contain in the stability regions S G1 and S G4 of controllers G1 and G 4 , respectively.

Hence these controllers transcript with controller G 3 . That is, assuming either G1 or G 4 as a

child controller G d , the trajectories resulting from the following equation,

x f (x) , x(td ) x dp , where

f ( x) a( x)  b( x) [ue  G d ( x  xe )]

will have a Jacobian that is asymptotically stable. These parent and child trajectories for various

decision points are presented in Figure 8. The ascent and descent modes are due to the

trajectories generated by child controllers G1 and G 4 respectively in SG1 and SG4 .

Fig. 8. Child FCMs bifurcating from parent FCM at decision points.

25
Fig. 9. Controller options influencing transcriptions.

Fig. 10. Initial conditions influencing transcriptions.

26
As it has been observed in the example, parent and child controller selection procedures are

context based. That is, they primarily depend on controller type (eigenvector variations) and

initial conditions. For instance, the transcription criterion shown in Figure 7 is modified as

shown in Figure 9, when controller G1 is replaced with controller G 2 . Here the transcriptions

discussed similar to Eqs. (23) are preserved for all decision points across the transients and

steady states. In contrast, these controllers in Figure 9 modify the transcriptions as shown in

Figure 10, when initial conditions in their respective stability regions are varied. Thus these

techniques in a way search whether decision points reside in the stability regions of the

controllers. Given an active controller as a parent, child controllers compatible to this parent may

be designed following the procedures presented in this paper. Generally, the transcription criteria

for decision points across transients of the initial conditions are unpredictable. Hence decision

points are favored at steady states.

Lastly, the control transcriptions are presented. The thrust coefficient ( CT (t ) ) in the aircraft of

Langelaan [17] must be positive. In thrust management, generally the transients are ignored but

the steady state values are programmable. Linearized model based nonlinear aircraft control

presented in this paper offers these steady state values for thrust management. Although positive

thrust management schemes for FCM analysis is important, in this paper, parent and child FCMs

with positive transients and steady states in thrust coefficient are considered to illustrate the

control transcriptions. For one of the trustable UAV transcriptions in Figure 9, the control

transcriptions are depicted in Figure 11 and 12, respectively. The control inputs for child FCMs

are presented in black color. The parent control inputs before bifurcations are presented in blue

color. In these problems, it is difficult to impose control input constraints.

27
Fig. 11. Control transcription with thrust coefficient.

Fig. 12. Control transcription with elevator inputs.

28
6. Conclusions

There has been a significant interest to automate the air defense system by using higher level

control architectures where the unmanned air vehicles are directed to accomplish their mission

operations. These decisions to engage an unmanned air vehicle require instantaneous controller

reconfiguration at inner loop so that it bifurcates from the intended path to the assigned task.

While the intended path is referred as the parent flight control mode, the path adopted to perform

the assigned task is referred as a child flight control mode. The decision points where the tasks

are assigned must reside in a region where stability regions of the controllers intersect. Given a

class of linear functional controllers for eigenvector variations, this paper establishes a search

technique for controller selection so that the decision points of the higher level control

architecture are naturally adaptive to stable reconfigurations of the controllers in their respective

stability regions. An unmanned air vehicle exhibiting these characteristics meets a transcription

criterion to define a trustable unmanned air vehicle for higher level control architectures. A three

degree of freedom aircraft is considered to illustrate the trustable unmanned air vehicle.

ACKNOWLEDMENT:

This research received no specific grant from any funding agency in the public, commercial, or

not-for-profit sectors.

References:

[1] J. Bacon, AirAsia flight 8501 climbed too fast before crash.USA Today, 9.11 PM EST,

January 20, 2015.

29
[2] C. Amato, G.D. Konidaris, G. Cruz, C.A. Maynor, J.P. How, L.P. Kaelbling, Planning for

decentralized control of multiple robots under uncertainty,” arXiv preprint

arXiv:1402.2871, 2014.

[3] W. Linda, S. Kannan, S. Sander, M. Guler, B. Heck, J.V.R. Prasad, D. Schrage, G.

Vachtsevanos, An open platform for reconfigurable control. IEEE Control Systems

Magazine, June 2001, pp. 49-63.

[4] J.C. Doyle, M. Csete, Architecture, constraints, and behavior. Proceedings of the

National Academy of Sciences, 108 (supplement 3), (2011), pp. 15624–15630.

[5] B.M. Muir, Trust between humans and machines, and the design of decision aids.

International Journal of Man-Machine Studies, 27(5), (1987), pp. 527–539.

[6] D.J. Leith, W.E. Leithhead, Appropriate realisation of gain scheduled controllers with
application to wind turbine regulation. International Journal of Control, 65, (1996), pp.
223-248.
[7] W.J. Rugh, Analytical framework for gain scheduling. IEEE Control Systems Magazine,
11 (1991), pp. 79-84.
[8] J.S. Shamma, M. Athans, Analysis of gain scheduled control for nonlinear plants. IEEE
Transactions on Automatic Control, 35, (1990), pp. 898-907.
[9] W.J. Rugh, J.S. Shamma, Research on gain scheduling. Automatica, 36, (2000), pp.
1401-1425.
[10] C.D.C. Jones, M.H. Lowenberg, T.S. Richardson, Tailored dynamic gains scheduled
control. Journal of Guidance, Controls, and Dynamics, 29(6), (2006), pp. 1271–1281.
[11] Y. Hou, Q. Wang, C. Dong, Gain scheduled control: Switched polytopic system
approach. Journal of Guidance, Control, and Dynamics, 34(2), (2011), pp.623-629.
[12] O.J. Oaks Jr., G. Cook, Piecewise linear control of nonlinear systems. IEEE Transactions

on Industrial Electronics and Control Instrumentation, IECI-23, (1976), pp. 56-63.

30
[13] R. Saeks, J. Murray, Fractional representation of algebraic geometry and the

simultaneous stabilization problem. IEEE Transactions on Automatic Control, 27, (1982),

pp. 895-903.

[14] M. Vidyasagar, N. Viswanadham, Algebraic design techniques for reliable stabilization.

IEEE Transactions on Automatic Control, 27, (1981), pp.1085-1095.

[15] V. Blondel, M. Gevers, Simultaneous stabilizability of three linear systems is rationally

undecidable. Mathematics of Control, Signals, and Systems, 6, (1993), pp.135-145.

[16] V. Blondel, Simultaneous Stabilization of Linear Systems, Lecture Notes in Control and

Information Sciences, 191, Spriger-Verlag, (1994). 184 pages.

[17] J.W. Langelaan, Gust energy extraction for mini and micro uninhabited aerial vehicles.

Journal of Guidance, Control, and Dynamics, 32(2), (2009), pp. 464-473.

[18] K. Sobel, F.J. Lallman, Eigenstructure assignment for the control of highly augmented

aircraft. Journal of Guidance, Control, and Dynamics, 12 (3), (1989), pp. 318-324.

[19] Y. Kim, D. Kum, J. Junkins, Design of small gain controller via iterative eigenstucture

assignment. AIAA Guidance, Navigation, and Control Conference, AIAA-93-3708-CP.

[20] C.R. Ashokkumar, W.P.G. York, Trajectory transcriptions for potential autonomy

features in UAV maneuvers. In: The proceedings of the AIAA Guidance, Navigation, and

Control Conference, San Diego, California, January 2016. AIAA-2016-0380.

[21] C.R. Ashokkumar, W.P.G. York, UAV control and simulation using trajectory

transcriptions. In: The Proceedings of the AIAA Modeling and Simulation Technologies,

San Diego, California, January 2016. AIAA-2016-0944.

31
[22] C.R. Ashokkumar, W.P.G. York, Observer based controllers for UAV maneuver options.

In: The Proceedings of the AIAA Guidance, Navigation, and Control Conference, San

Diego, California, January 2016. AIAA-2016-0643.

[23] J. Kautsky, N.K. Nichols, P. Van Dooren, Robust pole assignment in linear state

feedback. International Journal of Control, 41, (1985), pp. 1129–1155.

[24] B.D. Anderson, J.B. Moore, Optimal Control: Linear Quadratic Methods, Prentice Hall,

New Jersey, 1990.

[25] Y. Gu, B. Seanor, G. Campa, M.R. Napolitano, L. Rowe, S. Gururajan, S. Wan, Design

and flight testing evaluation of formation control laws. IEEE Transactions on Control

Systems Technology, 14(6), (2006), pp.1105-112.

[26] J.J.E. Slotine, W. Li, Applied Nonlinear Control, Prentice Hall International Inc., New

Jersey 07458, 1991, pp. 83-86.

[27] H.K. Khalil, Nonlinear Systems, Prentice Hall Inc., New Jersey 07458, 1996, p. 156.

[28] S. Barnett, Matrices in Control Theory, Van Nostrand Reinhold Company, London,

1985, p 85.

[29] R.A. Horn, C.R. Johnson, Matrix Analysis, Cambridge University Press, 1985, pp. 365.

[30] C.R. Ashokkumar, R.K. Yedavalli, Eigenstructure perturbations analysis in disjointed

domains for linear uncertain systems. International Journal of Control, 67(6), (1997), pp.

887-899.

32

You might also like