119
EE 580 — Linear Control Systems I. Models of Dynamical Systems Department of Electrical Engineering Pennsylvania State University Fall 2010 c 2010 Ji-Woong Lee 1.1 In troduction We are concerned with the analysis and synthesis of dynamical systems.  Analysis  of a dynamical system involves obtaining its mathematical model and then qualifying and quantifying its behavior by solv ing or simula ting the mathematical model . On the other hand,  synthesis  of a dynamical system inv olv es inte rconn ectin g several subsy stems such that the ov erall syste m meet s desir ed stability and performance properties. The behavior of a  dynamical system  is described by either dierential equations or dierence equati ons . A key featur e of such a sys tem is that its current beha vio r depends not only on the current input to the system but also on the past inputs to, and the initial condition of, the system. For example, if a mathematical description of a system is given by a dierential equation ˙ y(t) = u(t), t t 0 , or a dierence equation y(t + 1) = y (t) + u(t), t =  t 0 , t 0  + 1, ..., where  y(t) is the response of the system at time  t,  u(t) the input to the system at time t, and  t 0 the initial time, then the solution to such an equation reads y(t) =   t t 0 u(τ ) dτ  + y(t 0 ), t t 0 , or y(t) = t1 t 0 u(τ ) + y(t 0 ), t =  t 0  + 1, t 0  + 2, .... Clearly, the system response y (t) is determined by the past inputs  u(τ ),  τ < t. On the other hand, a trivial example of dynamical systems would be y (t) = u(t) 2 , where the system response  y (t) does not depend on the past inputs. Such a syste m b elong s to a special type of dynamical systems called static  systems, which are of less interest. To understand the concept of synthesis, consider the simple problem of charging a capacitor by a cur ren t source as depict ed in Fig. 1.1. Given the initi al time t 0  and nal time  t f  > t 0 , our objective is to charge the capacitor such that  v(t) =  v f  for all  t   t f  no matter what the initial 1

ee580_notes.pdf

  • Upload
    nagesh

  • View
    220

  • Download
    0

Embed Size (px)

Citation preview

Page 1: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 1/119

EE 580 — Linear Control Systems

I. Models of Dynamical Systems

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

c 2010 Ji-Woong Lee

1.1 Introduction

We are concerned with the analysis and synthesis of dynamical systems.   Analysis  of a dynamicalsystem involves obtaining its mathematical model and then qualifying and quantifying its behaviorby solving or simulating the mathematical model. On the other hand,  synthesis   of a dynamicalsystem involves interconnecting several subsystems such that the overall system meets desiredstability and performance properties.

The behavior of a   dynamical system   is described by either differential equations or differenceequations. A key feature of such a system is that its current behavior depends not only on thecurrent input to the system but also on the past inputs to, and the initial condition of, the system.For example, if a mathematical description of a system is given by a differential equation

y(t) = u(t), t ≥ t0,

or a difference equation

y(t + 1) = y(t) + u(t), t =  t0, t0 + 1, . . . ,

where  y(t) is the response of the system at time   t,   u(t) the input to the system at time   t, and   t0the initial time, then the solution to such an equation reads

y(t) =

   tt0

u(τ ) dτ  + y(t0), t ≥ t0,

or

y(t) =t−1

t0

u(τ ) + y(t0), t =  t0 + 1, t0 + 2, . . . .

Clearly, the system response y(t) is determined by the past inputs  u(τ ),  τ < t. On the other hand,a trivial example of dynamical systems would be  y(t) = u(t)2, where the system response  y(t) doesnot depend on the past inputs. Such a system belongs to a special type of dynamical systems calledstatic  systems, which are of less interest.

To understand the concept of synthesis, consider the simple problem of charging a capacitorby a current source as depicted in Fig. 1.1. Given the initial time   t0  and final time   tf   > t0, ourobjective is to charge the capacitor such that   v(t) =  vf   for all   t  ≥   tf   no matter what the initial

1

Page 2: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 2/119

EE 580 — Linear Control Systems    2

i (t)   v (t)

+

{

Figure 1.1: An RC circuit.

voltage  v(t0) across the capacitor is. This is a  control prob-lem , where the RC circuit is the  plant   (or  controlled sys-tem ), the current  i(t) is the control input  to the plant, andthe capacitor voltage  v(t) is the  controlled output  from theplant. Since   C v(t) =   i(t) for all   t  ≥   t0, any control sig-

nal  i  satisfying  C 

vf − v(t0)

 = tf 

t0 i(τ ) dτ   and i(t) = 0 fort  ≥   tf   will solve this control problem. In particular, thecontrol signal  i  given by

i(t) =

vf −v(t0)tf −t0

if  t ∈ [t0, tf );

0 if  t ≥ tf 

will do. This control scheme is an example of   open-loop control , where the entire control signal   i

is determined at the initial time based on the initial condition of the plant. However, open-loopcontrol is not practical. A slight error in the initial condition  v(t0) or a slight modeling error (e.g.,error in the value of capacitance   C ) will result in  v(tf )  =  vf  under open-loop control. Moreover,even if   v(tf ) =  vf   and   i(t) = 0 for   t  ≥   tf , a slight disturbance entering the circuit at some timet ≥ tf   will make  v(t) deviate from the desired voltage level  vf   and stay deviated.

A practical approach, on the other hand, is to employ  closed-loop control   (or feedback control ),where the control input at each time instant is corrected based on sensor measurements. Forexample, if measurements of the capacitor voltage   v(t) are available online, then a closed-loopcontrol that solves our problem of charging the capacitor is given by

i(t) =

vf −v(t)tf −t

  ,   if  t ∈ [t0, tf );

K 0

vf  − v(t)

  if  t ≥ tf ,(1.1)

where K 0  is any positive constant. It is readily seen that this closed-loop control scheme is insensi-tive to the initial condition or modeling error. Moreover, even if the system response  v(t) deviates

from the desired value  vf  due to a disturbance at some time t ≥ tf , the voltage v(t) will return to vf 

under this feedback control scheme. A feedback control scheme is implemented by interconnectingthe plant with another dynamical system, called the   controller ; the output of the controller is thecontrol input to the plant, and the   measured output   from the plant, as well as the   reference input 

v (t)RC

circuitv f 

i (t)+

{

K (t)

plantcontroller 

Figure 1.2: The closed-loop system formed by the RCcircuit in Fig. 1.1 and the controller in (1.1).

(i.e., the desired value for the controlledoutput of the plant), is the input to thecontroller. In the RC circuit example, the feedback control law  in (1.1) defines the con-troller in Fig. 1.2, where  v   is the measuredoutput from the plant (which happens to bethe same as the controlled output) and   vf 

the reference input. This controller hap-pens to be a static system represented bythe time-varying gain

K (t) =

  C tf −t

  if  t ∈ [t0, tf );

K 0   if  t ≥ tf .

The RC circuit example also illustrates two objectives of feedback control. One objective isthat of  performance , where the  closed-loop system   (or  feedback control system ) shown in Fig. 1.2 is

Page 3: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 3/119

EE 580 — Linear Control Systems    3

required to reach a target output voltage vf  at the terminal time  tf . It turns out that the feedbackcontrol given by (1.1) not only drives the capacitor voltage from an initial value  v(t0) to the finalvalue  v(tf ) = vf  but also minimizes the total heat energy dissipated by the resistor

J (i) =    tf 

t0

Ri(τ )2 dτ 

in doing so. In this sense, this controller achieves an optimal performance. (Design of optimalcontrollers is a topic of EE 581 and beyond the scope of EE 580.) The other control objective isthat of  stability , where the closed-loop system in Fig. 1.2 is required to maintain the output voltageat, or near, the reference value   v(t) =   vf   for all time   t  ≥   tf   subject to disturbances. Under thefeedback control law (1.1) with   K 0   >   0, we have that   i(t)   <   0 whenever   v(t)   > vf   and   i(t)   >  0whenever  v(t)  < vf , and so the stability objective is fulfilled. The larger the feedback gain  K 0   is,the more stable the closed-loop system becomes.

1.2 Notation

The set of real numbers is denoted by  R

, and the set of integers by  Z

. Given positive integersm  and   n, the Euclidean   n-space (i.e., the set of all   n-tuples (x1, . . . , xn) of real numbers   x1, . . . ,xn  ∈  R) is denoted by  Rn, and the set of all linear functions that map  R

n into   Rm is by  Rm×n.Elements of   Rn and   R

m×n are often represented by   n-dimensional column vectors and   m-by-n

matrices, respectively: If 

y1 = a11x1 + · · · + a1nxn,

...

ym = am1x1 + · · · + amnxn,

then we write

y =  Ax,

where

y =

y1...

ym

=

y1   · · ·   ym

T∈ R

m,   x =

x1...

xn

=

x1   · · ·   xn

T∈ R

n,

and

A =

a11   · · ·   a1n...

  . . .  ...

am1   · · ·   amn

=

a11   · · ·   am1...

  . . .  ...

a1n   · · ·   amn

T

∈ Rm×n.

More generally, if 

y1 = f 1(x1, . . . , xn),

...

ym = f m(x1, . . . , xn),

where  f 1, . . . ,  f m :  Rn → R  are given functions, then we write

y =  f (x)

Page 4: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 4/119

EE 580 — Linear Control Systems    4

with  y  = [y1   · · ·   ym]T,  x  = [x1   · · ·   xn]T, and  f  = [f 1,   · · ·   f m]T :  Rn → Rm understood.

If  x  is an  Rn-valued signal (i.e., vector-valued function of time) in the continuous-time domain,then we write  x :  R→ R

n. Similarly, if  x  is an  Rn-valued signal in the discrete-time domain, thenx :  Z  →   R

n. We will often use a single symbol, say,   x   to denote both a vector (in which casex  ∈  Rn) and a signal (in which case  x(t)  ∈  Rn for each   t): which is meant should be clear from

context. To emphasize that x   is a function of time, however, we may say that  x(·) is a signal, orthat x(t),  t ∈ R, (or x(t),  t ∈ Z) is a signal. Note that it is common, but incorrect, to say that x(t)is a signal;  x(t) is the value of signal  x  at time  t.

1.3 Mathematical Descriptions of Dynamical Systems

1.3.1 An Inverted Pendulum [1, Exercise 1.13]

 L

m

.

 M    ( )t 

Figure 1.3: An inverted pendulum.

An inverted pendulum consists of a pendulum of mass   m

mounted on a moving cart of mass  M  as shown in Fig. 1.3;the distance between the pivot and the center of gravity of the pendulum is L. The gravitational acceleration is  g, themoment of inertia with respect to the center of gravity   J ,and the damping factor associated with the cart   D. Theequation of motion for the inverted pendulum subject tothe external force  µ(t),  t  ≥  t0, applied to the cart is givenby the second-order ordinary differential equations

(J  + mL2)θ(t) = mgL sin θ(t)− mLρ(t)cos θ(t) (1.2a)

and

(M  + m)ρ(t) + D ρ(t)

= µ(t) + mLθ(t)2 sin θ(t)−mLθ(t)cos θ(t),   (1.2b)

where  θ(t) denotes the angular displacement of the pendulum from the vertical at time  t, and ρ(t)the displacement of the cart from the origin at time   t. Given the initial values  θ(t0),  θ(t0),   ρ(t0),and ρ(t0), and given the input signal µ(t), t ≥ t0, solving (1.2) yields the output values  θ(t) and ρ(t)for all  t ≥ t0.

1.3.2 Input-Output Representation of Dynamical Systems

 y Dynamical

system

u

Figure 1.4: A dynamical system.

Suppose that we have a dynamical system that generatesan output signal  y(t),   t  ≥  t0, according to an input signalu(t),   t  ≥  t0, as depicted in Fig. 1.4, where  t0   is the initialtime at which the system begins to operate. An exampleis the inverted pendulum described in Section 1.3.1: Withzero initial values, there exist functions  h1 and  h2 such that

θ(t) = h1(µ(s), t0  ≤ s ≤ t),

ρ(t) = h2(µ(s), t0  ≤ s ≤ t)

for all  t ≥ t0.

Page 5: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 5/119

EE 580 — Linear Control Systems    5

In general, if a system of ordinary differential equations has  m  input variables  u1(t), . . . ,  um(t)and   l  output variables   y1(t), . . . ,  yl(t), then, with “zero initial values,” there is a function  h  suchthat

y(t) = h(t, u(s), t0  ≤ s ≤ t),   (1.3)

where u(s) is the m-tuple (u1(s), . . . , um(s)) of input variables, and y(t) the l-tuple (y1(t), . . . , yl(t))

of output variables. Here, the output cannot be related to the input via  y(t) =  h(u(t)) for any(memoryless) function   h :  Rm →   R

l. A dynamical system has an internal memory; that is, theoutput  y(t) for each time instant   t   is a function of the input signal   u(s),   s  ≥   t0. For instance,the inverted pendulum in Section 1.3.1 is a dynamical system with one input (i.e.,   m   = 1) andtwo outputs (i.e.,   l  = 2). Denoting the entire signals  u(t),   t  ≥   t0, and  y(t),   t  ≥   t0, by  u  and  y,respectively, the  input-output representation  (1.3) is often written

y =  H (u) or   y =  H u   (1.4)

where  H  is an operator (i.e., signal-valued function of signals) defined by a function  h  as in (1.3).Note that, according to (1.3), the output at time   t   does not depend on the input   u(s) for anys > t; that is, the present output is determined solely by the past and present inputs. A dynamicalsystem possessing this property is said to be  causal .

1.3.3 State-Space Representation of Dynamical Systems

Consider the inverted pendulum described in Section 1.3.1. Let  x1(t) = θ(t),  x2(t) =  θ(t),  x3(t) =ρ(t), and  x4(t) = ρ(t); let   u(t) =  µ(t),   y1(t) =  θ (t), and  y2(t) =  ρ(t). Then it is readily seen thatthere exist functions f 1, . . . ,  f 4  such that

xi(t) = f i(x1(t), x2(t), x3(t), x4(t), u(t)), i = 1, . . . ,   4,   (1.5)

for all  t ≥ t0, and functions  g1  and  g2  such that

yi(t) = gi(x1(t), x2(t), x3(t), x4(t), u(t)), i = 1,   2,   (1.6)

for all   t  ≥   t0; in this particular case, we have x1(t) =   x2(t), x3(t) =   x4(t),   y1(t) =   x1(t), andy2(t) =   x3(t). Given initial values   x1(t0), . . . ,   x4(t0), and given an input signal   u(t),   t   ≥   t0,equations (1.5) and (1.6) define the output variables  y1(t) and  y2(t) for all  t ≥ t0.

In general, a system of high-order ordinary differential equations can be reduced to a system of first-order ordinary differential equations: If the former has n   initial values   x1(t0), . . . ,  xn(t0),  m

input variables  u1(t), . . . ,  um(t), and  l  output variables  y1(t), . . . ,  yl(t), then there exist functionsf :  R× R

n ×Rm → R

n and  g :  R× Rn × R

m → Rl such that the latter is of the form

x(t) = f (t, x(t), u(t)),   (1.7a)

y(t) = g(t, x(t), u(t)),   (1.7b)

where  x(t) = [x1(t)   · · ·   xn(t)]T,  x(t) = [x1(t)   · · ·  xn(t)]T,  u(t) = [u1(t)   · · ·   um(t)]T, and  y(t) =[y1(t)   · · ·   yl(t)]T for all   t  ≥   t0. Equations (1.7) comprise the   state-space representation   (or  state model ) of a dynamical system that can be described by a system of ordinary differential equations;in particular, equation (1.7a) is called the   state equation , and (1.7b) the   output equation . Thevectors  x(t),  u(t), and  y(t) are called the state ,  input , and  output , respectively.

The number n of state variables is called the order  of the system (1.7). For instance, the invertedpendulum described by two second-order ordinary differential equations results in a fourth-order

Page 6: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 6/119

EE 580 — Linear Control Systems    6

system represented by (1.5) and (1.6). Since  n,   m, and   l  are all finite, the system (1.7) is oftensaid to be  finite-dimensional . In particular, since  m,   l  ≥  1, the system is called a  multiple-input multiple-output (MIMO)  system; the special case of  m  =  l  = 1 is called a single-input single-output (SISO)  system. If the function  f , called the  vector field , is independent of  u  (i.e., if there exists afunction  f   such that   f (t, x, u) =  f (t, x) for all   t  ∈  R,  x  ∈  Rn, and  u  ∈  Rm), then the first-order

ordinary differential equation (1.7a) is said to be  homogeneous   (or unforced ); otherwise, it is calledinhomogeneous   (or   forced ).

1.3.4 Importance of State Concept

According to J. C. Willems [2], “the point of view which most control theorists adhere to is that,while actual systems are most naturally viewed in an input/output setting, when it comes tocomputations state models have superior properties.” One of the most important among suchproperties is what we call “finite-memory property.” Consider the state-space representation (1.7) of a dynamical system. The output equation (1.7b) describes the output  y(t) in terms of the state x(t)(as well as the input  u(t)) via a “memoryless” function  g . This means that the “internal memory”of a dynamical system at time   t   is absorbed in the state  x(t) at time   t. Moreover, according to

the state equation (1.7a), the number of state variables remains finite and constant throughout theoperation of the plant; that is, the “size” of the internal memory of a finite-dimensional dynamicalsystem remains finite and constant.

The finite-memory property of the state-space model is particularly useful in the context of feedback controller design. At each time instant   t, a feedback controller generates the input  u(t)based on the information available to it up to time  t; if the output signal is measured online, thenthis information consists of  y(s) for all   s  ∈   [t0, t] and   u(s) for all   s  ∈   [t0, t). However, it is well-known that for many control objectives this information can be replaced with  x(t) whenever thestate is perfectly observed at time  t  (i.e., whenever y(t) = x(t)). Another advantage of state-spacemethods is that, unlike input-output models, state-space models provide a full description of theinternal behavior of dynamical systems. For this reason, state-space models are called   internal 

descriptions   of dynamical systems. On the other hand, input-output models describe the systembehavior under zero initial conditions and are not necessarily able to capture the evolution of internal state variables that are not controlled or observed. Thus input-output models are calledexternal descriptions  of dynamical systems. We will study these aspects of state-space approachesin more detail during this course.

1.4 Classification of Dynamical Systems

1.4.1 A Target Maneuver Model for Tracking Radar [3, Example V.B.1]

Let us consider a target maneuver model that tracking radar systems often use. With a sam-pling period   T , a digital computer samples the position of the moving target as a point in two-

dimensional Cartesian coordinate. Then the target position at sampling instant  tT   is given by thepair (ρ1(t), ρ2(t)) for all  t  =  t0,  t0 + 1, . . . , where  t0  ∈ Z defines the initial sampling instant  t0T . If the acceleration of the target at sampling instant  tT   is (µ1(t), µ2(t)), then Newton’s law says thatthe velocity (ν 1(t), ν 2(t)) of the target at sampling instant  tT  approximately satisfies

ρi(t + 1) = ρi(t) + T ν i(t),

ν i(t + 1) = ν i(t) + T µi(t)

Page 7: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 7/119

EE 580 — Linear Control Systems    7

for all  t =  t0,  t0 + 1, . . . . This leads to the state equation

x1(t + 1)x2(t + 1)x3(t + 1)x4(t + 1)

=

1 0   T    00 1 0   T 

0 0 1 00 0 0 1

x1(t)x2(t)x3(t)x4(t)

+

0 00 0T    00   T 

u1(t)u2(t)

  (1.8a)

and output equation

y1(t)y2(t)

 =

1 0 0 00 1 0 0

x1(t)x2(t)x3(t)x4(t)

(1.8b)

where   x1(t) =   ρ1(t),   x2(t) =   ρ2(t),   x3(t) =   ν 1(t),   x4(t) =   ν 2(t),   u1(t) =   µ1(t),   u2(t) =   µ2(t),y1(t) = ρ1(t), and  y2(t) = ρ2(t) for all  t  =  t0,  t0 + 1, . . . . (In a target model like this, the unknownacceleration inputs  µ1(t) and  µ2(t) are often considered as “noise.”)

Just like other dynamical systems, this target maneuver model has many different state-spacerepresentations. Let   x1(t) and   x2(t) be as above, but put   x3(t) =   ρ1(t − 1) − T 2µ1(t − 1) and

x4(t) = ρ2(t − 1)− T 2µ2(t − 1). Then the state equation now reads

x1(t + 1)x2(t + 1)x3(t + 1)x4(t + 1)

=

2 0   −1 00 2 0   −11 0 0 00 1 0 0

x1(t)x2(t)x3(t)x4(t)

+

0 00 0

−T 2 00   −T 2

u1(t)u2(t)

for all  t =  t0,  t0 + 1, . . . .

1.4.2 Continuous-Time and Discrete-Time Systems

A dynamical system represented by ordinary differential equations as in (1.7) is called a  continuous-

time   system. For example, the inverted pendulum model described in Section 1.3.1 leads to acontinuous-time system. On the other hand, a   discrete-time  system has the state-space represen-tation consisting of first-order difference equations of the form

x(t + 1) = f (t, x(t), u(t)),   (1.9a)

y(t) = g(t, x(t), u(t)) (1.9b)

for t  =  t0,  t0 + 1, . . . , where t0  ∈ Z  is the initial time. If  u(t) and y(t) are the the  m-tuple of inputvariables and   l-tuple of output variables, respectively, then the input-output representation of adiscrete-time system is written as in (1.3); denoting the entire sequences  u(t0),  u(t0 + 1), . . . , andy(t0),  y(t0 + 1), . . . , by u  and  y, respectively, we may write as in (1.4), where   H   is an operatordefined by   h. Examples of discrete-time systems include the state-space representations of the

target maneuver model described in Section 1.4.1.

1.4.3 Time-Varying and Time-Invariant Systems

The system of ordinary differential equations (1.7) or that of difference equations (1.9) is said to betime-invariant   (or  autonomous ) if the functions  f   and  g  are independent of  t (i.e., if there exist  f 

and g  such that  f (t, x, u) =  f (x, u) and  g(t, x, u) = g(x, u) for all   t ∈ R, x ∈ Rn, and  u ∈ Rm).Time-invariance is characterized by the shift-invariance  of the input-output representation (un-

der zero initial conditions); that is, a time shift in the input signal leads to the same time shift

Page 8: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 8/119

EE 580 — Linear Control Systems    8

t t 

u t ( )

u t ( )

 y t ( )

 y t ( )

 H 

 H 

Figure 1.5: A time-invariant system  H .

in the output signal as depicted in Fig. 1.5.The inverted pendulum model in Sec-tion 1.3.1 and the target maneuver modelin Section 1.4.1 both lead to time-invariantsystems. A dynamical system which is

not time-invariant is called   time-varying .An example of time-varying systems is theclosed-loop system in Fig. 1.2; since the con-troller gain   K (t) is a function of time, theoverall system is time-varying.

1.4.4 Linear and Nonlinear Systems

A continuous-time system (1.7) or a discrete-time system (1.9) is said to be  linear   if the functions f 

and  g  are of the form

f (t, x, u) = A(t)x + B(t)u,

g(t, x, u) = C(t)x + D(t)u

for all   t   ∈   R,   x   ∈   Rn, and   u   ∈   R

m, where   A(t)   ∈   Rn×n,   B(t)   ∈   R

n×m,   C(t)   ∈   Rl×n, and

D(t) ∈ Rl×m are matrices determined by  f (t, ·, ·) and g(t, ·, ·). In particular, a linear time-invariant 

(LTI)  system has

f (t, x, u) = Ax + Bu,

g(t, x, u) = Cx + Du

for all   t  ∈   R,   x  ∈   Rn, and  u  ∈  Rm, where  A,   B,   C, and   D  are constant matrices. If a systemis linear but not time-invariant, then it is called a   linear time-varying (LTV)  system. If a systemis not linear, it is called   nonlinear . The inverted pendulum example in Section 1.3.1 leads to a

nonlinear system. On the other hand, the target models in Section 1.4.1 are LTI systems, and theoverall system in Fig. 1.2, where the reference input  vf   is considered an input to the system, is anLTV system.

In general, linearity is characterized by the  superposition principle , which says that a dynamicalsystem is linear if and only if its input-output representation   H  as in (1.4) satisfies the followingproperty under zero initial conditions: Whenever  y =  H u  and  z =  H v, we have

αy + β z =  H (αu + β v)

for all  α,  β  ∈ R.

References[1] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[2] J. C. Willems, “The concept of a dynamical system in control,” in Proceedings of the 33rd IEEE Conference on Decision and Control , vol. 3, 1994, p. 2099.

[3] H. V. Poor,   An Introduction to Signal Detection and Estimation , 2nd ed. New York, NY:Springer, 1994.

Page 9: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 9/119

EE 580 — Linear Control Systems

II. Solution and Linearization of State-Space Models

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

2.1 Notation and Definitions

The   Euclidean (vector) norm  x   of   x ∈   Rn is defined by x   =

√ xTx   =

ni=1 x2

i

1/2. The

following properties of a vector norm can be easily checked:  x  = 0 if and only if  x  = 0 ∈  Rn;

αx = |α|x for all x ∈ R

n

and α ∈ R; and the triangle inequality  x + y ≤ x + y holds forx, y ∈ R

n. Moreover, the Cauchy-Schwarz inequality  says that |xTy| ≤ xy for all x, y ∈ Rn [1,

Lemma 5.4.13 & Eq. (5.4.15)]. This vector norm ·  on  Rn induces a matrix norm on  Rm×n: Thespectral norm   (or operator norm ) of  A ∈ R

m×n is defined by

A = max{Ax : x = 1} = max{Ax/x :  x = 0}.

For example, if  A  = [ 1 00 2 ], then A = 2 because, over all  x  = [x1 x2]T ∈ R

2 with x = 

x21 + x2

2

= 1, the norm Ax  = 

x21 + 4x2

2   is maximized by  x  = [0 1]T. Although we will use the samesymbol  ·   to denote both the vector norm and matrix norm, it should be clear from contextwhich is meant. Similarly to the Euclidean vector norm, the spectral norm satisfies the followingproperties:

 A

 = 0 if and only if  A =  0

 ∈ R

m×n;

 αA

 =

 |α

|A

 for all  A

 ∈ R

m×n and  α

 ∈ R;

and A + B ≤ A + B   for  A,  B ∈  Rm×n. Moreover, the spectral norm is  submultiplicative ;that is, BA ≤ BA   and Ax ≤ Ax   for all   B ∈   R

l×m,   A ∈   Rm×n, and   x ∈   R

n [1,Theorem 5.6.2]. The identity matrix   I   (i.e., the matrix representation of the identity map) hasunity spectral norm: I = 1.

For r > 0 and x0 ∈ Rn, the set B(r, x0) = {x ∈ R

n : x−x0 < r} is called an open ball   centeredat x0  and of radius  r. If  S  ⊂ R

n, a point  x ∈ S  is said to be an  interior point   of  S  whenever thereexists some  r > 0 such that  B (r, x) ⊂ S ; otherwise  x  is said to be a  boundary point   of  S . The setof all interior points (resp. boundary points) of  S  is called the interior   (resp.   boundary ) of  S . Theset  S   is said to be  open   if the interior of  S   is  S   itself;  S   is said to be  closed   if its complement isopen. That is, an open set does not contain any of its boundary points, and a closed set containsall its boundary points. Simplest examples include open intervals (a, b), closed intervals [a, b], and

intervals like (a, b] and [a, b) that are neither open nor closed. By convention, the empty set  ∅ andthe entire set  Rn are both open and closed. A set S  ⊂ R

n is said to be bounded  if there exists somer >  0 such that  S  ⊂  B(r, 0) (or equivalently, x  < r   for all  x ∈  S ); if  S   is closed and bounded,then  S   is said to be  compact .

A sequence (x1, x2, . . . ) of vectors   xk ∈   Rn is said to  converge   to a vector   x ∈  R

n, which iscalled the   limit   of the sequence (xk), if for any given  ε >  0 there exists a nonnegative integer  K such that xk − x  < ε   for all  k ≥  K ; in this case, we write either   xk →  x  or limk→∞ xk   =  x.For example, while the sequence (xk) with   xk   =   k/(1 + k) is convergent and has   xk →   1, thesequence with  xk  = 1 + (−1)k is not convergent. A sequence (f 1, f 2, . . . ) of functions  f k :  D → R

m

1

Page 10: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 10/119

EE 580 — Linear Control Systems    2

is said to converge (pointwise)  to a function f  :  D → Rm if the sequence (f 1(x), f 2(x), . . . ) converges

to  f (x) for all  x ∈ D. The sequence (f k) is said to  converge uniformly  to a function  f   if for everyε >   0 there exists a nonnegative integer   K   such that f k(x) − f (x)   < ε   whenever   k ≥   K   andx ∈   D; note that   K   does not depend on   x   in the case of uniform convergence. Geometrically,uniform convergence means that when   k > K   the entire graph of   f k   must lie within the strip

bounded by the graphs of  f k − ε  and  f k +  ε. For example, let  f k,  gk,  f  : (0, 1) →R

 be defined byf k(x) = max{−x + 1/k, 0},  gk(x) = max{−kx + 1, 0}, and  f (x) = 0 for  x ∈  (0, 1); while  gk →  f uniformly, we have  f k → f  only pointwise.

A function f  :  D → Rm, where D ⊂ R

n, is said to be continuous at the point  x0 ∈ D  if for everyε > 0 there exists a  δ > 0 such that f (x) − f (x0) < ε  whenever x − x0 < δ  and  x ∈ D. If thefunction  f   is continuous at  x0, then we write limx→x0 f (x) = f (x0). The function f   is said to becontinuous (on  D)   if it is continuous at all  x0 ∈  D. If for every  ε >  0 there exists a  δ >  0 suchthat f (x) − f (x0)  < ε   whenever x − x0  < δ   and  x,  x0 ∈  D, then   f   is said to be  uniformly continuous   on  D; note that  δ  does not depend on  x  or  x0   in the case of uniform continuity. Forexample, if  f (x) = 2x   for  x ∈  R

n, then  f   is uniformly continuous on  Rn because, for any  ε >  0,we have f (x) − f (x0) < ε  whenever x − x0 < ε/2; on the other hand, if  f (x) = x2, then  f is continuous on  Rn but not uniformly so because, for any  δ > 0, choosing an  x0  with

 x0

= 1/δ 

and setting  x  = (1 + δ 2/2)x0  gives x − x0 = δ/2 < δ  but

|f (x) − f (x0)| = |xTx − xT0 x0| =

(x + x0)T(x − x0) =

(2 + δ 2/2)x0

T(δ 2/2)x0 > 1;

continuity of   f   follows from the fact that, for any   ε >   0 and   x0 ∈   Rn, inequality x − x0   < 

ε + x02 − x0   implies |f (x) − f (x0)| = (x + x0)T(x − x0) ≤ x + x0x − x0 < ε. It is a

fact that,  if  f  :  D → Rm is continuous on  D  and if  D   is compact, then  f   is uniformly continuous 

on  D   [2, Theorem 2.44] and bounded on  D   (i.e., there is some  M >  0 such that f (x) ≤  M   forall  x ∈ D) [2, Corollary 2.40].

The set of every continuous function that maps  D   into  Rm is denoted by  C (D, Rm). For anypositive integer  k, we define  C k(D,Rm) as the set of every function  f  ∈ C (D,Rm) such that

∂  jf l∂xi1

1 · · · ∂xinn

∈ C (D,R)

holds for i1 + · · ·+ in  = j , for j  = 1, . . . ,  k, and for l  = 1, . . . , m. If  f  ∈ C k(D,Rm), then f   is calleda C k-function . In particular, a C 1-function is said to be  continuously differentiable , meaning that itis differentiable and its derivative is continuous. The  Jacobian matrix   of a C 1-function f  :  D → R

m

at  x0 ∈ D, where D ⊂ Rn, is defined by

∂f 

∂ x(x0) =

∂f 1∂x1

(x0)   · · ·   ∂f 1∂xn

(x0)

...  . . .

  ...∂f m

∂x1 (x0)   · · ·  ∂f m

∂xn (x0)

∈ Rm×n.

2.2 Solutions of State Equations [3, Section 1.5][4, Section 1.10]

2.2.1 Existence of Solutions

Given a function  f  :  D → Rn, where D   is an open, nonempty, and connected subset of  R×R

n, andgiven initial data (t0, x0) ∈ D, consider the homogeneous state equation

x(t) = f (t, x(t)) (2.1a)

Page 11: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 11/119

EE 580 — Linear Control Systems    3

subject tox(t0) = x0; (2.1b)

equivalently, consider the integral equation

x(t) = x0 +    t

t0

f (s, x(s)) ds,

where    tt0

f (s, x(s)) ds =

 tt0

f 1(s, x(s)) ds... t

t0f n(s, x(s)) ds

 =

 tt0

f 1(s, x1(s), . . . , xn(s)) ds... t

t0f n(s, x1(s), . . . , xn(s)) ds

.

A function  φ :  J  → Rn, where J   is an open interval in  R, is said to be a solution  of the initial-value

problem (2.1) on the interval  J  if it satisfies the following conditions:

•   φ ∈ C 1(J,Rn);

•   (t,φ(t)) ∈ D  for all  t ∈ J ;

•  φ(t) = f (t,φ(t)) for all  t ∈ J ;

•   φ(t0) = x0.

The following result shows that the continuity of  f   is sufficient for the existence of a solutionto the initial-value problem. Moreover, it shows that one can solve the initial-value problem bothforward and backward in time.

Theorem 2.1   If  f  ∈  C (D, Rn)  and   (t0, x0) ∈ D, then the initial-value problem   (2.1)  has at least one solution on the interval  (t0 − c, t0 + c)  for some  c > 0.

2.2.2 Uniqueness of Solutions

For example, consider the initial-value problem

x(t) = x(t)1/3, x(0) = 0.   (2.2)

Applying Theorem 2.1 to  f (t, x) = x1/3 yields that there exists a solution to this problem. Indeed,a solution is given by

φ(t) =

0 if  t < 0;

(2t/3)3/2 if  t ≥ 0.(2.3a)

However, it is not unique sinceφ(t) = 0, t ∈ R,   (2.3b)

is also a solution. This example shows that continuity of   f   is not enough for the initial-value

problem (2.1) to have a unique solution.A function f  :  D → Rn is said to satisfy a  Lipschitz condition  (or to be  Lipschitz continuous   in

the second argument) if, for every compact subset K  of  D, there exists a constant  LK  > 0 (calleda  Lipschitz constant  with respect to the second argument) such that

f (t, x) − f (t, y) ≤ LK x − yfor all (t, x), (t, y)  ∈   K . Lipschitz continuity is a stronger condition than continuity, and issufficient for the uniqueness of solutions to initial-value problems. We again assume that thedomain  D ⊂ R×R

n of the vector field  f   is open, nonempty, and connected.

Page 12: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 12/119

EE 580 — Linear Control Systems    4

Theorem 2.2   If  f  ∈ C (D, Rn)  and  (t0, x0) ∈ D, and if  f  satisfies a Lipschitz condition, then the initial-value problem   (2.1)  has at most one solution on the interval  (t0 − c, t0 + c)   for any  c >  0.

2.2.3 Continuation of Solutions

Once a solution to the initial-value problem has b een found on a (possibly small) time interval con-

taining the initial time t0, the process of extending it to a larger time interval is called continuationof a solution.

As an example, consider the initial-value problem (2.2), and suppose one has obtained a solutionφ(t) = 0 on the interval (−1, 0). In this case, φ   is continuable in more than one way: Both (2.3a)and (2.3b) are extensions of this  φ  and are solutions to (2.2) on the entire time interval  R. As asecond example, consider the initial-value problem

x(t) = x(t)2, x(0) = 1,

which has a solution  φ(t) = (1 − t)−1 on (−1, 1). Clearly, this solution is continuable to the left(i.e., backward in time) to −∞ but not to the right (i.e., forward in time).

Theorem 2.3   If  f  ∈

C (D,Rn), where  D  =  J ×Rn  for some open interval  J 

 ⊂R, and if  f  satisfies 

a Lipschitz condition, then for any  (t0, x0) ∈ D the initial-value problem  (2.1) has a unique solution that exists on the entire interval  J .

This continuation result will be used to show that the initial-value problem with a linear,possibly time-varying state equation has a unique solution on the entire time interval if the matrixcoefficients are continuous, or piecewise continuous, in time.

2.2.4 Successive Approximations of Solutions

If the initial-value problem (2.1) has a unique solution  φ, a classical method of constructing ap-proximations to   φ   is the   method of successive approximations   (or   Picard iterations ). Given afunction  f 

 ∈ C (D,Rm), where  D

 ⊂ R

×Rn is open, nonempty, and connected, and given initial

data (t0, x0) ∈  D,  successive approximations   for (2.1) are the sequence (φ0,φ1, . . . ) of functionsφm  given by

φ0(t) = x0,   (2.4a)

φm+1(t) = x0 +

   tt0

f (s,φm(s)) ds, m = 0,   1,   2, . . . ,   (2.4b)

for all  t   in some interval  J  ⊂ R containing  t0.

Theorem 2.4   If  f  ∈ C (D, Rn)  and  (t0, x0) ∈ D, and if  f  satisfies a Lipschitz condition, then the successive approximations  φm   given by   (2.4)   exist on the interval   (t0 − c, t0 +  c)   for some   c >  0,are continuous on   (t0 − c, t0 +  c), and converge uniformly, as  m → ∞, to the unique solution of 

(2.1)  on  (t0 − c, t0 + c).

2.3 Linearization of State-Space Models

2.3.1 Linearized Equations [3, Section 1.6]

Consider the state-space model

x(t) = f (t, x(t), u(t)),   (2.5a)

y(t) = g(t, x(t), u(t)),   (2.5b)

Page 13: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 13/119

EE 580 — Linear Control Systems    5

where  f , g :  R× D1 × D2 → Rn,  D1 ⊂ R

n, and  D2 ⊂ Rm. Let  φ ∈ C (R, D1) and  ψ ∈ C (R, D2) be

given signals. If  f , g ∈ C 1(R× D1 × D2,Rn) and if  φ  and  ψ  satisfy

φ(t) = f (t,φ(t),ψ(t))

for all t

∈R, then we can linearize the state-space model (2.5) about (φ,ψ) in the following manner.

For each   x ∈   C (R, D1) and   u ∈   C (R, D2) satisfying the state equation (2.5a) for all   t ∈   R,define

δ x(t) = x(t) −φ(t), δ u(t) = u(t) −ψ(t) (2.6a)

for all  t ∈ R. If 

∂f 

∂ x =

∂f 1∂x1

· · ·   ∂f 1∂xn

...  . . .

  ...∂f n∂x1

· · ·   ∂f n∂xn

,

  ∂f 

∂ u =

∂f 1∂u1

· · ·   ∂f 1∂um

...  . . .

  ...∂f n∂u1

· · ·   ∂f n∂um

are the Jacobian matrices of  f  with respect to the second and third arguments, respectively, thenwe obtain

˙δ x(t) = f (t, x(t), u(t)) − f (t,φ(t),ψ(t))

= f (t, δ x(t) + φ(t), u(t)) − f (t,φ(t),ψ(t)) + f (t,φ(t), δ u(t) + ψ(t)) − f (t,φ(t), u(t))   =0

= ∂f 

∂ x(t,φ(t),ψ(t)) δ x(t) +

 ∂f 

∂ u(t,φ(t),ψ(t)) δ u(t) +  F 1(t, δ x(t), u(t)) +  F 2(t, δ u(t)),

where

F 1(t, δ x(t), u(t)) = f (t, δ x(t) + φ(t), u(t)) − f (t,φ(t), u(t)) −  ∂f 

∂ x(t,φ(t),ψ(t)) δ x(t),

F 2(t, δ u(t)) = f (t,φ(t), δ u(t) + ψ(t))−

f (t,φ(t),ψ(t))−

 ∂f 

∂ u(t,φ(t),ψ(t)) δ u(t).

Fix a  t ∈ R. What we want to have is that, as  δ x(t) → 0 and  δ u(t) → 0, the functions  F 1  and  F 2vanish “faster” than  δ x(t) and  δ u(t), respectively. Standard terminology to denote such functionsis  o(·), the  “little o” function : If a function  h :  D →  R

m, with  D  being a subset of  Rn containingthe origin, is such that limx→0 h(x)/x = 0, then h  is said to be  o(x). With this terminology, itcan be shown that  F 1(t, δ x(t), u(t)) is o(δ x(t)) for each  t ∈ R and  u(t) ∈ R

m, and that  F 2(t, δ u(t))is  o(δ u(t)) for each  t ∈ R. Letting

A(t) =  ∂f 

∂ x(t,φ(t),ψ(t)) and   B(t) =

 ∂f 

∂ u(t,φ(t),ψ(t)),   (2.6b)

we obtain

˙δ x =  A(t) δ x(t) + B(t) δ u(t) + o(δ x(t)) + o(δ u(t))

for each  t ∈ R. Similarly, letting

δ y(t) = g(t, x(t), u(t)) − g(t,φ(t),ψ(t)),   (2.6c)

C(t) =  ∂g

∂ x(t,φ(t),ψ(t)),   and   D(t) =

  ∂g

∂ u(t,φ(t),ψ(t)) (2.6d)

yieldsδ y(t) = C(t) δ x(t) + D(t) δ u(t) +  o(δ x(t)) + o(δ u(t))

Page 14: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 14/119

EE 580 — Linear Control Systems    6

for each  t ∈ R. As δ x(t) → 0 and δ u(t) → 0, we have

˙δ x(t) = A(t) δ x(t) + B(t) δ u(t),

δ y(t) = C(t) δ x(t) + D(t) δ u(t)(2.7)

for each  t∈R. The state-space model (2.7) is called the   linearization  of (2.5) about the solution  φ

and the input signal  ψ.If the state-space model (2.5) is autonomous (i.e., time-invariant) and homogeneous (i.e., un-

forced), so thatx(t) = f (x(t)) and   y(t) = g(x(t)),

and if  φ   is a constant solution of it, so that  φ(t) =  x0   for all   t ∈   R   for some   equilibrium point x0 ∈ R

n satisfying f (x0) = 0, then its linearization (2.5) reads

˙δ x(t) = A δ x(t) and   δ y(t) = C δ x(t),

where  A  and  C are constant matrices.Discrete-time state equations of the form  x(t + 1) =  f (t, x(t), u(t)) and  y(t) =  g(t, x(t), u(t))

can be linearized in a virtually identical manner to the continuous-time case, and their linearizedmodels are of the form

δ x(t + 1) = A(t) δ x(t) + B(t) δ u(t),

δ y(t) = C(t) δ x(t) + D(t) δ u(t),

where  δ x,  δ u,  δ y,  A(t), B(t),  C(t), and  D(t) are as in (2.6).

2.3.2 Special Case: Taylor’s Theorem [2, Theorem 3.11][5, Theorem 28.2]

Let  f  : [a, b] → R  and its first  k  derivatives be continuous on [a, b] and differentiable on (a, b), andlet  x0

 ∈  [a, b]. Then for each  x

 ∈  [a, b] with  x

 =  x0   there exists a point  c   between  x  and  x0   such

that

f (x) = f (x0) + ∂f 

∂x(x0)(x − x0) +

  1

2!

∂ 2f 

∂x2(x0)(x − x0)2 + · · ·

+  1

k!

∂ kf 

∂xk(x0)(x − x0)k +

  1

(k + 1)!

∂ k+1f 

∂xk+1(c)(x − x0)k+1.   (2.8)

Here it is clear that all the terms on the right-hand side of equality (2.8), except for the first two,are  o(x − x0); that is, as  x → x0, we have  f (x) → f (x0) +   ∂f 

∂x (x0)(x − x0). Moreover, if  f (x0) = 0(i.e., if  x0   is an equilibrium of x =  f (x)), then we have a linearization of  f  about x0, which is givenby   ∂f 

∂x (x0)δx.

2.3.3 Example: Inverted Pendulum

The motion of an inverted pendulum is described by the ordinary differential equations

(J  +  mL2)θ(t) = mgL sin θ(t) − mLρ(t)cos θ(t),   (2.9a)

(M  +  m)ρ(t) + D ρ(t) = µ(t) + mLθ(t)2 sin θ(t) − mLθ(t)cos θ(t),   (2.9b)

Page 15: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 15/119

EE 580 — Linear Control Systems    7

where θ(t) denotes the angular displacement of the pendulum from the vertical at time  t, and ρ(t)the displacement of the cart from the origin at time   t. A state-space model of (2.9), with theargument t  omitted for simplicity, reads

x1  =  x2,

x2  =

 (M  +  m)mgL sin x1

−mLu cos x1

−m2L2x2

2 sin x1 cos x1 + mLDx4 cos x1

(M  + m)(J  +  mL2) − m2L2 cos2 x1 ,

x3  =  x4,

x4  = (J  +  mL2)u + mL(J  +  mL2)x2

2 sin x1 − m2gL2 sin x1 cos x1 − D(J  +  mL2)x4

(M  + m)(J  +  mL2) − m2L2 cos2 x1

andy1 =  x1, y2 =  x3

with state variables x1 =  θ, x2 =  θ, x3 =  ρ, and x4 = ρ; input variable  u  =  µ; and output variablesy1 =  θ  and  y2 =  ρ. Using the procedure in Section 2.3.1, we linearize this model about the solution(φ1(t), φ2(t), φ3(t), φ4(t)) = (0, 0, v0t, v0),  t ∈ R, and the input ψ(t) = Dv0,  t ∈ R, as

˙δ x(t) = A δ x(t) + B δ u(t),δ y(t) = C δ x(t),

where, regardless of the constant  v0,

A =  1

(M  + m)J  +  mM L2

0 (M  +  m)J  +  mM L2 0 0mgL(M  + m) 0 0   MLD

0 0 0 (M  + m)J  +  mM L2

−m2gL2 0 0   −D(J  +  mL2)

,

B =  1

(M  + m)J  +  mM L2

0

−mL0

J  +  mL2

,   C =

1 0 0 00 0 1 0

.

2.3.4 Importance of Linearized Equations

A real-world process or environment can have several different models of varying levels of complexity,and linear equations may arise in a natural manner from the simplest possible model. However,more often than not, linear equations result from the process of linearizing nonlinear equations.Linear systems are important because of the following:

•  Unlike nonlinear systems, linear systems are very well understood, and there exists a relativelycomplete theory for analyzing and synthesizing linear systems;

•  Linearized equations may give us qualitative information about the underlying, often very

complicated, nonlinear behavior.

2.4 Solutions of Linear State Equations

2.4.1 Existence and Uniqueness [3, Section 1.7]

Given matrices  A(t) ∈ Rn×n,  B(t) ∈ R

n×m for  t ∈ R, and given a function  u :  R → Rm and initial

data (t0, x0) ∈ R× Rn, consider the inhomogeneous linear state equation

x(t) = A(t)x(t) + B(t)u(t) (2.10a)

Page 16: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 16/119

EE 580 — Linear Control Systems    8

with the initial statex(t0) = x0.   (2.10b)

Theorem 2.5   Suppose that   A ∈   C (J,Rn×n),   B ∈   C (J,Rn×m), and   u ∈   C (J,Rm), where   J   is some open interval in   R. Then for any   t0 ∈  J   and any  x0 ∈  R

n, the initial-value problem   (2.10)has a unique solution on the entire interval  J .

Proof . Since  A(t) and  B(t)u(t) are continuous in  t, the function  f (t, x) = A(t)x + B(t)u(t) iscontinuous in (t, x). Moreover, for any closed interval J 0 ⊂ J  there is a Lipschitz constant  L0 ≥ 0such that

f (t, x) − f (t, y) = A(t)(x − y) ≤ A(t)x − y ≤ L0x − yfor all (t, x), (t, y) ∈ J 0 ×R

n; here, L0  is taken to be maxt∈J 0 A(t), which exists as J 0   is compactand A(·) is continuous on  J 0  [2, Corollary 2.40]. (Here, continuity of  A(·) follows from that of  ·  and  A(·).) Therefore, by Theorem 2.3, there exists a unique solution to (2.10) on  J 0. Sincethis argument holds for any closed interval  J 0 ⊂ J , a unique solution exists on  J .  

A function defined on an interval   J   of   R   is said to be  piecewise continuous   if it has at most

a finite number of discontinuities over any bounded interval in   J . Theorem 2.5 extends to thecase where   A(·),   B(·), and  u(·) are piecewise continuous because one can solve the initial-valueproblem (2.10) piece by piece. For example, if  A(·), B(·), and u(·) are continuous on [t0, tf ) exceptpossibly at  t1, . . . ,   tN −1 ∈  R  with  t0  < t1  < · · · < tN −1  < tf , then we obtain a solution  φ(t) overall  t ∈ [t0, tf ) by solving  N   initial-value problems as follows:

x(t) = A(t)x(t), t ∈ [t0, t1);   x(t0) = x0   ⇒   φ(t), t ∈ [t0, t1);

x(t) = A(t)x(t), t ∈ [t1, t2);   x(t1) = limτ →t−1

φ(τ )   ⇒   φ(t), t ∈ [t1, t2);

...

x(t) = A(t)x(t), t ∈ [tN −1, tf );   x(tN −1) = limτ →t−N −1

φ(τ )   ⇒   φ(t), t ∈ [tN −1, tf ).

2.4.2 Peano-Baker Series [3, Section 1.8]

Application of Theorem 2.4 to (2.10) leads to the following result:

Theorem 2.6   Suppose that  A ∈ C (R,Rn×n),  B ∈ C (R,Rn×m), and  u ∈ C (R,Rm). Then for any t0 ∈ R and any  x0 ∈ R

n, the initial-value problem   (2.10)  has the unique, continuously differentiable solution 

x(t)  total

solution

=  Φ(t, t0)x0   homogeneous

solution

+

   tt0

Φ(t, s)B(s)u(s) ds   particular solution

 for all  t∈R, where the  state transition matrix  Φ(t, t0)  is given by the  Peano-Baker series

Φ(t, t0) = I +

   tt0

A(s1) ds1 +

   tt0

A(s1)

   s1t0

A(s2) ds2ds1

+

   tt0

A(s1)

   s1t0

A(s2)

   s2t0

A(s3) ds3ds2ds1 + · · ·   ,

which converges uniformly on any bounded interval in  R   containing   t  and   t0.

Page 17: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 17/119

EE 580 — Linear Control Systems    9

In the time-invariant case whereA(t) = A  for all  t, the Peano-Baker series simplifies to

Φ(t, t0) =

∞k=0

Ak(t − t0)k

k!  = I + A(t − t0) +

 1

2A2(t − t0)2 +

 1

6A3(t − t0)3 + · · ·

with the convention that 0! = 1 and  A0

=   I; in this case, we write  Φ(t, t0) =  eA(t−t0)

, which iscalled the  matrix exponential .

2.4.3 Discrete-Time Case

Consider the initial-value problem determined by discrete-time linear systems of the form

x(t + 1) = A(t)x(t) + B(t)u(t),   (2.11a)

x(t0) = x0.   (2.11b)

Define the discrete-time state transition matrix by

Φ(t, t0) = I   if  t  =  t0;A(t − 1) · · · A(t0) if  t > t0.

Then the unique solution to (2.11) is given by

x(t)  total 

solution 

=   Φ(t, t0)x0   homogeneous

solution 

+t−1s=t0

Φ(t, s + 1)B(s)u(s)

   particular solution 

for all   t  =  t0,   t0 + 1, . . . . (Note that, unlike the continuous-time case, solving (2.11) backward intime for t < t0  is in general not possible.) In the time-invariant case where  A(t) = A  for all t

∈R,

we haveΦ(t, t0) = At−t0 .

References

[1] R. A. Horn and C. R. Johnson, Matrix Analysis . Cambridge, UK: Cambridge University Press,1985.

[2] C. C. Pugh, Real Mathematical Analysis . New York, NY: Springer, 2002.

[3] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[4] ——, Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

[5] S. R. Lay, Analysis with an Introduction to Proof , 2nd ed. Englewood Cliffs, NJ: Prentice Hall,1990.

Page 18: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 18/119

EE 580 — Linear Control Systems

III. Input-Output Description of Linear Systems

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

3.1 Definitions and Facts About Matrix Inversion

A function f  :  Rn →  Rm is said to be one-to-one (or injective) if  f (x) = f (x) whenever x  = x; it is

said to be onto (or surjective) if, for every  y ∈  Rm, there exists an x ∈  R

n such that f (x) = y. If  f   is

both one-to-one and onto, then it is called a one-to-one correspondence (or bijection). Bijections areinvertible functions; that is, if  f  is bijective, then there exists a unique  f −1 such that f −1(f (x)) = xfor all  x  ∈  R

n and such that  f (f −1(y)) = y  for all  y  ∈  Rm.

A matrix A  ∈  Rn×n is said to be nonsingular (or invertible) if there exists a matrix A−1 ∈  R

n×n

called the inverse of   A   such that   A−1A   =   I. The inverse   A−1 is unique whenever it exists,and satisfies  AA−1 =   I   as well. A nonsingular matrix is a representation of a linear one-to-onecorrespondence. If   A,   B   ∈   R

n×n are both nonsingular, then it is readily seen that (AB)−1 =B−1A−1. The following are useful facts for computing the inverse of a matrix:

•   Inverse of a Partitioned Matrix [1, Appendix A.22]. Let   A   ∈   Rn×n be nonsingular and

partitioned as

A = A11   A12

A21   A22with  A11  ∈  R

n1×n1 ,  A22  ∈  Rn2×n2 , and  n1 + n2  = n. If  A−1

11   exists, then  A−1 is partitionedas

A−1 =

A−1

11  + A−1

11 A12∆−1

11 A21A−1

11  −A−1

11 A12∆−1

11

−∆−111

 A21A−111

  ∆−111

with

∆11 = A22 − A21A−111 A12.

If  A−1

22  exists as well, then the (1, 1) block of  A−1 can be replaced with  ∆−1

22  where  ∆22  =

A11 − A12A−122

 A21; ∆11  (resp.   ∆22) is known as the  Schur complement   of  A11  (resp.   A22).

•  A Matrix Inversion Formula [2, Section 0.7.4]. Let  B ∈  R

n×n

be nonsingular and given by

B =  A + XRY

for some  A  ∈  Rn×n, X ∈  R

n×m,  R ∈  Rm×m, and  Y ∈  R

m×n. Then

B−1 = A−1 − A−1X

R−1 + YA−1X−1

YA−1

whenever  A and  R  are nonsingular.

1

Page 19: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 19/119

EE 580 — Linear Control Systems    2

3.2 Impulse Response of Linear Systems [3, Section 2.4][4, Section 1.16]

3.2.1 Impulse Function

© k (t)

0{1/k    1/k t

Figure 3.1: Functions  φk.

Define a sequence (φk) of functions  φk   ∈  C (R,R),  k  = 1,2, .. . , by

φk(t) =k(1 − k|t|) if  |t| ≤ 1/k;

0 otherwise.(3.1)

Although the sequence (φk) does not converge to any func-tion, it satisfies the following:

Lemma 3.1   Let   φk   ∈   C (R,R),   k   = 1,   2, . . . , b e a s  

in   (3.1). Then 

limk→∞

   ∞

−∞

φk(t − s)f (s) ds =  f (t)

 for all  f  ∈ C (R,R)  and for all   t ∈  R.

Proof . Choose an  f  ∈ C (R,R) and a t  ∈  R. Since  f  is continuous, the minimum (resp. maximum)of  f (s) over all  s  ∈  [t − 1/k,t + 1/k] is achieved at some point  αk  (resp. β k) in [t − 1/k,t + 1/k] forany positive  k  ∈  Z. It follows from 

  ∞

−∞

φk(t − s) ds =

   t+1/kt−1/k

φk(t − s) ds = 1

that

f (αk) ≤    ∞

−∞

φk(t − s)f (s) ds ≤  f (β k)

for each positive integer  k. Since  αk   and  β k  both converge to   t  as  k  → ∞  and since  f   is contin-uous at   t, we conclude that limk→∞ f (αk) = limk→∞ f (β k) =  f (t) and hence that the sequence  

−∞φk(t − s)f (s) ds

 must converge to  f (t) as  k  → ∞.  

Based on Lemma 3.1, we write

f (t) =

   ∞

−∞

δ (t − s)f (s) ds   (3.2)

or simplyf (t) = (δ  ∗ f )(t)

for   f   ∈   C (R,R) and   t   ∈   R. Equation (3.2) is a symbolic representation, not an integral, forthe following reasons: First of all, (φk) does not converge to a function; secondly, even if (gk) isa sequence of continuous functions converging to a function   g, interchange of limit and integral(i.e., the equality lim

  gk   =

   lim gk) is not guaranteed unless the convergence is uniform [5,

Corollary 4.7]. The symbol  δ   is called the   Dirac delta function   (or  unit impulse ). In particular, if f  ∈ C ([0, ∞),R), then

f (t) =

   ∞

0−δ (t − s)f (s) ds

for all  t  ≥  0. (The lower limit of this integration starts from 0− to include the case of  t = 0.)

Page 20: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 20/119

EE 580 — Linear Control Systems    3

3.2.2 Integral Representation of Linear Systems

A SISO linear operator  H :  C (R,R) →  C (R,R) is said to admit an   integral representation   if thereexists an integrable function  h :  R × R →  R  such that

(Hu)(t) =    ∞

−∞

h(t, s)u(s) ds

for all u  ∈  C (R,R) and for all t  ∈  R; the function h is called the kernel  of the integral representationof   H . For example, the differential equation y(t) =   u(t),   t  ≥   t0, with the zero initial conditiony(t0) = 0, admits the integral representation  y(t) = (Hu)(t) =

 ∞

−∞h(t, s)u(s) ds, where its kernel

is given by  h(t, s) = 1 for  s  ∈   [t0, t] and  h(t, s) = 0 otherwise. If  φk   ∈  C (R,R) is as in (3.1) foreach  k, then Lemma 3.1 leads to

limk→∞

   ∞

−∞

h(t, τ )φk(s − τ ) dτ  =

   ∞

−∞

h(t, τ )δ (s − τ ) dτ  = h(t, s)

for all   t,  s  ∈  R. Thus we identify the kernel h  with the   impulse response :   h(t, s) is the output attime  t  due to the unit impulse applied at the input at time  s.

Now consider a general MIMO linear operator H : C (R,Rm

) →  C (R,Rl

). The linear operator H is said to admit an integral representation if there exists an integrable function  H :  R × R →  R

l×m

(i.e., a function  H whose entry  hij :  R × R →  R  in row i  and column  j  is integrable for all  i  and  j )such that

H(t, s) =

h11(t, s)   · · ·   h1m(t, s)...

  . . .  ...

hl1(t, s)   · · ·   hlm(t, s)

for  t,  s ∈  R and such that

(H u)(t) =

   ∞

−∞

H(t, s)u(s) ds   (3.3)

−∞ h11(t, s)u1(s) + · · · + h1m(t, s)um(s) ds... ∞

−∞

hl1(t, s)u1(s) + · · · + hlm(t, s)um(s)

ds

for all  u  = [u1   · · ·   um]T ∈  C (R,Rm) and for all   t  ∈  R. If we let all components of  u(s) be zeroexcept for the  jth component, then the  ith component of the output (H u)(t) assumes the form

yi(t) =

   ∞

−∞

hij(t, s)u j(s) ds.

Thus  hij(t, s), which is entry (i, j) of  H(t, s), represents the  ith component of the output at time tdue to a unit impulse applied at the  j th component of the input at time  s  (while all other compo-

nents of the input are identically zero), and the matrix  H(t, s) is called the impulse response matrix of the system  H .

3.2.3 Causality and Relaxedness

It is readily seen that a MIMO linear system H  is causal if and only if the impulse response matrixsatisfies H(t, s) = 0  whenever  t < s. For causal systems, (3.3) reduces to

(H u)(t) =

   t−∞

H(t, s)u(s) ds.

Page 21: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 21/119

EE 580 — Linear Control Systems    4

Causality implies that the system output does not depend on the future input, and is a propertythat every physical system possesses.

As C.-T. Chen pointed out in [6], “in developing the input-output description, before an inputis applied, the system must be assumed to be relaxed or at rest, and that the output is excitedsolely and uniquely by the input applied thereafter.” We have assumed so far that the system is

at rest at  t =  −∞: a MIMO linear system  H   is said to be   relaxed   (or   at rest ) at  t =  t0   if we have(H u)(t) = 0 for  t ≥  t0   whenever  u(t) = 0 for  t ≥  t0. If the system  H   is causal and if it is relaxedat  t  =  t0, then we may write

(H u)(t) =

   tt0

H(t, s)u(s) ds.   (3.4)

3.2.4 Time-Invariance

If a MIMO linear system  H   is time-invariant, then we have  H(t, s) =  H(t − s, 0) for all   t,  s  ∈  R,and so one often abuse notation and write (3.3) as

(H u)(t) =    ∞

−∞

H(t − s)u(s)ds.

The right-hand side of this equality is called the  convolution integral   of  H  and  u  and is written as

(H u)(t) = (H ∗ u)(t).

It is easily seen that a linear time-invariant system is causal if and only if  H(τ ) = 0 for all  τ < 0.

3.2.5 Discrete-Time Systems

The   discrete-time impulse function (or  unit pulse )  δ  is defined by

δ (t) = 1 if  t  = 0;

0 otherwise

for all  t  ∈  Z, and we may write

u(t) =∞

s=−∞

δ (t − s)u(s)

for all  u :  Z  →  R  and for all   t  ∈   Z. Suppose that a discrete-time SISO linear operator   H   has arepresentation of the form

(Hu)(t) =∞

s=−∞

h(t, s)u(s) (3.5)

when u :  Z →  R and  h :  Z × Z →  R  are such that the infinite sum on the right-hand side of (3.5) iswell defined for all t ∈  Z. For example, the sum is guaranteed to be finite if either of the followingholds [3, Pages 60–61]:

• 

s=−∞ h(t, s)2 < ∞  for each   t ∈  Z, and 

s=−∞ u(s)2 < ∞(due to the Schwarz inequality);

• 

s=−∞ |h(t, s)| <  ∞  for each  t ∈  Z, and sups∈Z |u(s)| <  ∞,where sups∈Z |u(s)|  denotes the supremum (i.e., least upper bound) on  |u(s)|  over all  s ∈  Z.

Page 22: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 22/119

EE 580 — Linear Control Systems    5

Then we have∞

s=−∞

h(t, τ )δ (s − τ ) = h(t, s)

for all   t,  s  ∈  Z, and so  h   is said to be the  discrete-time impulse response   (or   unit pulse response )of  H :   h(t, s) represents the response of  H  at time  t  to a unit pulse occurring at time  s.

If  H   is a discrete-time MIMO linear operator that admits an integral representation, then (3.5)is generalized to

(H u)(t) =

∞s=−∞

H(t, s)u(s) (3.6)

=

∞s=−∞

h11(t, s)u1(s) + · · · + h1m(t, s)um(s)

...

s=−∞

hl1(t, s)u1(s) + · · · + hlm(t, s)um(s)

for some  H :  Z × Z →  Rl×m and for all  u :  Z →  Rm such that the infinite sum is well defined. The

matrix H(t, s) is called the  discrete-time impulse response matrix , and the entries of  H(t, s) have asimilar interpretation to their continuous-time counterparts. The system  H   is causal if and only if H(t, s) = 0 whenever t < s. On the other hand, the system is  relaxed   at  t =  t0   if (H u)(t) = 0 fort ≥  t0   whenever  u(t) = 0 for  t ≥  t0. If  H   is causal and relaxed at  t0, then

(H u)(t) =t

s=t0

H(t, s)u(s).   (3.7)

Finally, if  H   is time-invariant, then  H (t, s) = H (t − s, 0); we abuse notation and write (3.6) as aconvolution sum 

(H u)(t) =∞

s=−∞

H(t − s)u(s).

3.3 Transfer Function Description of Linear Time-Invariant Systems

3.3.1 Continuous-Time Case

The   Laplace transform  of a function  f  : [0, ∞) →  R  (including the Dirac delta function) is denotedby  f (s) and defined as

f (s) =

   ∞

0−e−stf (t) dt

where s  is a complex variable. (The lower limit 0− is to include a possible impulse in f  at time 0.)Tables of Laplace transform pairs and properties are in [3, Page 92].

Suppose that a continuous-time MIMO linear time-invariant system  H  is causal and relaxed att = 0. Then taking the Laplace transform of both sides of its integral representation

y(t) =

   ∞

0

H(t − s)u(s) ds,

with  H(t − s) = 0 for  t < s, yieldsy(s) = H(s)u(s)

Page 23: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 23/119

EE 580 — Linear Control Systems    6

due to the convolution property of Laplace transforms, where

y(s) =

y1(s)...

yl(s)

,   u(s) =

u1(s)...

um(s)

,   H(s) =

h11(s)   · · ·   h1m(s)...

  . . .  ...

hl1(s)   · · ·   hlm(s)

.

The matrix H(s), which is the Laplace transform of the impulse response matrix, is called thetransfer function matrix  of the system  H .

3.3.2 Discrete-Time Case

The  z -transform  of a function  f  :  {0, 1, . . . } →  R is denoted by  f (z) and defined as

f (z) =∞t=0

z−tf (t).

Tables of  z -transform pairs and properties can be found in [3, Page 113].

If  H   is a discrete-time MIMO linear time-invariant system which is causal and relaxed at  t  = 0,then it can be shown that taking the  z-transform of both sides of its representation

y(t) =∞s=0

H(t − s)u(s),

with  H(t − s) = 0 for  t < s, leads to

y(z) = H(z)u(z)

where

y(z) = y1(z)

...yl(z) ,   u(z) =

u1(z)...

um(z) ,   H(z) = h11(z)   · · ·   h1m(z)

...  . . .

  ...

hl1(z)   · · ·   hlm(z) .

The matrix H(z) is called the   discrete-time transfer function matrix  of the system H .

3.4 Response of Linear Systems

3.4.1 Continuous-Time Case [3, Section 3.4]

Consider the continuous-time linear system of the form

x(t) = A(t)x(t) + B(t)u(t),   (3.8a)

y(t) = C(t)x(t) + D(t)u(t) (3.8b)

with the initial statex(t0) = x0.

If  A ∈  C (R,Rn×n),  B ∈  C (R,Rn×m),  C ∈  C (R,Rl×n),  D ∈  C (R,Rl×m), and  u ∈  C (R,Rm), thenthe   total system response   is given by

y(t) = C(t)Φ(t, t0)x0   zero-input 

response 

+

   tt0

C(t)Φ(t, s)B(s)u(s) ds + D(t)u(t)   zero-state response 

Page 24: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 24/119

EE 580 — Linear Control Systems    7

for all   t ∈  R, where  Φ(t, s) is the state transition matrix of the system (3.8). Setting  x0  = 0 andcomparing this with the integral representation (3.4) for causal linear systems relaxed at  t0  yieldsthe impulse response matrix

H(t, s) = C(t)Φ(t, s)B(s) + D(t)δ (t − s) if  t  ≥  s;

0   if  t < s.If the system (3.8) is time-invariant, so that  A(t) = A, B(t) = B, C(t) = C, and D(t) = D  for

all  t, then taking the Laplace transform of both sides of (3.8a) and (3.8b) with  t0 = 0 leads to

sx(s) − x(0) = Ax(s) + Bu(s),

y(s) = Cx(s) + Du(s),

where the first equality follows from the time differentiation property of Laplace transforms. There-fore, we obtain

y(s) = C(sI − A)−1x(0) + C(sI − A)−1Bu(s) + Du(s),   (3.9)

which, together with  x(0) = 0, yields the expression for the transfer function matrix: H(s) = C(sI − A)−1B + D.

Similarly, equation (3.9), together with  u(t) = 0 for all   t  (i.e.,  u(s) = 0), yields that the Laplacetransform of the state transition matrix  Φ(t, 0) =  eAt,   t  ≥  0, is given by the matrix (sI − A)−1,which is known as the   resolvent   of  A.

3.4.2 Discrete-Time Case [3, Section 3.5]

The   total system response  of the discrete-time linear system

x(t + 1) = A(t)x(t) + B(t)u(t),   (3.10a)

y(t) = C(t)x(t) + D(t)u(t) (3.10b)

with the initial statex(t0) = x0

is given by

y(t) = C(t)Φ(t, t0)x0   zero-input 

response 

+

t−1s=t0

C(t)Φ(t, s + 1)B(s)u(s) + D(t)u(t)   zero-state response 

for all  t =  t0,  t0 + 1, . . . . Comparing this with (3.7) yields that the discrete-time impulse responsematrix is as follows:

H(t, s) =C(t)Φ(t, s + 1)B(s) if  t > s;

D(t) if  t  =  s;

0   if  t < s.

If the system (3.10) is time-invariant with the initial time   t0  = 0, then taking the  z-transformof both sides of (3.10a) and (3.10b) with  t0 = 0 leads to

zx(z) − zx(0) = Ax(z) + Bu(z),

y(z) = Cx(z) + Du(z),

Page 25: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 25/119

EE 580 — Linear Control Systems    8

where the first equality follows from the time advance property of   z-transforms. Therefore, weobtain that the  z-transform of the total response  y(t),  t = 0, 1, . . . , is given by

y(z) = Cz(zI − A)−1x(0) + C(zI − A)−1Bu(z) + Du(z).

Putting  x(0) = 0 yields that the discrete-time transfer function matrix is expressed as H(z) = C(zI − A)−1B + D.

On the other hand, putting  u(t) = 0 for all   t   (i.e.,  u(z) = 0) yields that the  z-transform of thestate transition matrix  Φ(t, 0) = At,  t  = 0, 1, . . . , is given by  z(zI − A)−1.

3.4.3 Equivalent Descriptions of LTI Systems

Given a state-space model (A, B, C, D) of a linear time-invariant system of the form

x(t) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t),  or

x(t + 1) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t),

let us change the state vector  x  to a new state vector  x  via

x =  Px   ⇔

x1...

xn

 =

 p11x1 + · · · + p1nxn...

 pn1x1 + · · · + pnnxn

  ⇔   x =  P−1x

for some nonsingular matrix  P  = ( pij)  ∈  Rn×n, whose entry (i, j) is  pij   for all   i  and   j. Then we

obtain a new state-space model ( A, B, C, D) of the form

˙x(t) = Ax(t) + Bu(t),

y(t) =

 Cx(t) +

 Du(t),

orx(t + 1) = Ax(t) + Bu(t),

y(t) =

 Cx(t) +

 Du(t),

where A =  P−1AP, B =  P−1B, C =  CP, and D =  D. It follows from

y(s) = C(sI − A)−1x(0) + C(sI − A)−1Bu(s) + Du(s)

= C(sI − A)−1x(0) + C(sI − A)−1 Bu(s) + Du(s),

and similar equations for the discrete-time case, that the two state-space models (A, B, C, D)and (A, B, C, D) share the same external (input-output) descriptions. Therefore, they are calledequivalent   internal (state-space) descriptions of an LTI system.

3.5 Example: A Quarter-Car Model [7, Example 2.11]

Fig. 3.2 shows a simplified model of an automobile suspension system of one wheel. In this “quarter-car” model, M 1   is the mass of a quarter of the car body,  M 2  the mass of a wheel,  B1   the dampingfactor of the shock absorber,  B2   the damping factor of the tire,   K 1   the stiffness of the springs,and K 2  the stiffness of the tire. If  z1(t) and z2(t) denote the vertical displacements of the car bodyand the wheel from their nominal heights, respectively, at time  t, then the equations of motion forthe system are given by

M 1z1(t) = f 1(t) − B1

z1(t) −  z2(t)

− K 1

z1(t) − z2(t)

,

M 2z2(t) = f 2(t) − B1

z2(t) −  z1(t)

− K 1

z2(t) − z1(t)

− B2 z2(t) − K 2z2(t),

Page 26: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 26/119

EE 580 — Linear Control Systems    9

z 1(t )M 1

M 2

K 2

K 1   B 1

B 2

z 2(t )

f  1(t )

f  2(t )

Automobile

Suspensionsystem

Wheel

Tire

Figure 3.2: A quarter-car model.

where  f 1(t) is the force at time   t   due to aerodynamic load and load transfer effects, and  f 2(t) isthe force at time   t  due to the tire copying the road bumps. Under zero initial conditions, theseequations give

M 1s2 + B1s + K 1   −B1s − K 1−B1s − K 1   M 2s2 + (B1 + B2)s + (K 1 + K 2)

z1(s)z2(s)

 =f 1(s)

f 2(s)

or, equivalently,

z1(s)z2(s)

 =

M 1s2 + B1s + K 1   −B1s − K 1

−B1s − K 1   M 2s2 + (B1 + B2)s + (K 1 + K 2)

−1 f 1(s)

f 2(s)

=

  1

∆(s)

M 2s2 + (B1 + B2)s + (K 1 + K 2)   B1s + K 1

B1s + K 1   M 1s2 + B1s + K 1

f 1(s)

f 2(s)

,

where

∆(s) = M 1M 2s4 + (B1M 1 + B1M 2 + B2M 1)s3

+ (K 1M 1 + K 1M 2 + K 2M 1 + B1B2)s2 + (K 1B2 + K 2B1)s + K 1K 2.

Let y  =  z1  be the output, and let  u = [f 1 f 2]T be the input. Then the transfer function matrix forthe quarter-car system is given by

 H(s) =

M 2s2 + (B1 + B2)s + (K 1 + K 2)

∆(s)

B1s + K 1∆(s)

,

and we have the input-output description y(s) =

 H(s)u(s) in the Laplace-transform domain.

While there is a unique external description of the quarter-car model, there are infinitely manyinternal descriptions for the model. For instance, if we define x1  = z1,  x2  = z1,  x3  = z2,  x4  = z2,u1 = f 1, u2  =  f 2, and  y  =  z1, then the resulting state-space model reads

x(t) = Ax(t) + Bu(t);

y(t) = Cx(t) + Du(t),

Page 27: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 27/119

EE 580 — Linear Control Systems    10

where  x  = [x1  x2 x3 x4]T,  u = [u1 u2]T, and

A =

0 1 0 0−K 1/M 1   −B1/M 1   K 1/M 1   B1/M 1

0 0 0 1K 1/M 1   B1/M 2   −(K 1 + K 2)/M 2   −(B1 + B2)/M 2

,   B =

0 01/M 1   0

0 00 1/M 2

,

C = 1 0 0 0 ,   D = 0 0 .

Indeed, we recover the transfer function matrix via H(s) = C(sI − A)−1B + D.

References

[1] T. Kailath,  Linear Systems . Englewood Cliffs, NJ: Prentice-Hall, 1980.

[2] R. A. Horn and C. R. Johnson, Matrix Analysis . Cambridge, UK: Cambridge University Press,1985.

[3] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[4] ——,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

[5] C. C. Pugh,  Real Mathematical Analysis . New York, NY: Springer, 2002.

[6] C.-T. Chen,   Linear System Theory and Design , 2nd ed. New York, NY: Oxford UniversityPress, 1984.

[7] C. L. Phillips and R. D. Harbor,  Feedback Control Systems , 4th ed. Upper Saddle River, NJ:Prentice Hall, 2000.

Page 28: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 28/119

EE 580 — Linear Control Systems

IV. Finite-Dimensional Vector Spaces

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

4.1 Introduction

The space of complex numbers is denoted by  C. Depending on the context, we will use the  n-tuplerepresentation of the elements of  Cn and write  x = (x1, . . . , xn) ∈  C

n, or use their column vector

representations and write  x

 = [x1

  · · ·   xn]

T

∈ C

n

. Since R

 ⊂ C

, the same notational conventionapplies to the elements of the Euclidean space  Rn.Some of the basic linear algebraic tools are required for rigorous analysis of linear dynamical

systems. We will first study the properties of linear maps between finite-dimensional vector spaces.For our purposes, finite-dimensional vector spaces can be taken to be either  Rn or  Cn dependingon the context. This study will prove useful in the analysis and synthesis of linear systems.

4.2 Vector Spaces [1, Appendix A.1]

4.2.1 Fields

Definition 4.1   Let  F  be a set, and let  +  and   ·  be two binary operations that map  F  × F   into  F .The tuple  (F, +, ·), or the set  F   (with  +  and  ·  understood), is called a  field  if the following axioms are satisfied:

(a)   α + β  =  β  + α  and  αβ  =  βα   for all  α,  β  ∈ F ; (commutativity)α + (β  + γ ) = (α + β ) + γ   and  α(βγ ) = (αβ )γ   for all  α,  β ,  γ  ∈ F ; (associativity)α(β  + γ ) = αβ  + αγ   for all  α,  β ,  γ  ∈ F . (distributivity)

(b) There exists  0F   ∈ F   such that  0F  + α =  α  for all  α ∈ F ; (identity w.r.t.   +)There exists  1F   ∈ F   with  1F   = 0F , such that  1F α =  α  for all  α ∈ F . (identity w.r.t.   ·)

(c) For each  α ∈ F   there exists  −α ∈ F   such that  α + (−α) = 0F ; (inverse w.r.t.   +)For each  α ∈ F   with  α = 0F , there exists  α−1 ∈ F   such that  αα−1 = 1F . (inverse w.r.t.   ·)

If  (F, +, ·)  is a field, then  +  and   ·  are called   addition  and  multiplication, respectively.

The set of real numbers R and the set of complex numbers C, equipped with the usual additionand multiplication, are fields. the set of rational functions (i.e., rational fractions over polynomialswith real coefficients), denoted  R(s), is a field. The set {0, 1}  does not form a field if we use theusual definition of addition and multiplication, because the element 1 + 1 = 2 is not in the set{0, 1}. However, if we define 0 + 0 = 1 + 1 = 0, 1 + 0 = 0 + 1 = 1, and 1 · 0 = 0 · 1 = 0 · 0 = 0,1 · 1 = 1, then it can be verified that ({0, 1}, +, ·) is a field, called the field of binary numbers. Theset of integers   Z  and the set of polynomials with real coefficients do not form a field under theusual addition and multiplication because they have no multiplicative inverse.

1

Page 29: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 29/119

EE 580 — Linear Control Systems    2

4.2.2 Vector Spaces

Definition 4.2   Let  V  be a set, let  F  be a field, let  +  be a binary operation that maps  V  ×V   into V ,and let   ·  be a binary operation that maps  F  × V   into  V . The tuple   (V,F, +, ·), or the pair   (V, F )(with  +  and   ·  understood) or the set  V   (with  F ,  +, and   ·  understood), is called a  vector space  (or linear space) over  F  if the following axioms are satisfied:

(a)   x + y =  y  + x  for all  x,  y ∈ V  ;x + (y + z) = (x + y) + z   for all  x,  y,  z ∈ V  ;There exists a unique  0V    ∈ V   such that  0V    + x =  x  for all  x ∈ V . (zero vector or origin)

(b)   (αβ )x =  α(βx)  for all  α,  β  ∈ F  and for all  x ∈ V ;0F x = 0V    and  1F x =  x  for all  x ∈ V .

(c)   α(x + y) = αx + αy  for all  α ∈ F  and for all  x,  y ∈ V ;(α + β )x =  αx + βx  for all  α,  β  ∈ F  and for all  x ∈ V  .

If   (V,F, +, ·)   is a vector space, then the elements of   V    and   F   are called   vectors   and   scalars,respectively, and the binary operations  +  and  ·  are called  vector addition and  scalar multiplication.

If  F  is a field, denote the set of  n-tuples of elements of  F   by F n. With  x = (x1, . . . , xn) ∈ F n,y = (y1, . . . , yn) ∈ F n, and  α ∈ F , define vector addition and scalar multiplication by

x + y = (x1 + y1, . . . , xn + yn) and   αx = (αx1, . . . , α xn).   (4.1)

Then  F n is a vector space over the field  F . In particular,  Rn and  Cn are vector spaces, and calledthe  n-dimensional  real vector space   and  n-dimensional  complex vector space , respectively. Theseare examples of finite-dimensional vector spaces. If  x  = (x1, . . . , xn)  ∈  F n with either  F   =  R  orF   = C, we will use the column vector representation  x = [x1   · · ·   xn]T whenever necessary.

4.2.3 Linear Subspaces

Definition 4.3   A nonempty subset   W   of a vector space   V    over a field   F   is called a   (linear)subspace  of  V   if  αx + βy  ∈ W   whenever  α,  β  ∈ F   and  x,  y ∈ W .

A linear subspace is a vector space in its own right. If  W 1   and  W 2  are subspaces of a vectorspace  V , then  W 1 ∩ W 2   is also a linear subspace of  V  . To show that a set V   is a vector space, itsuffices to show that it is a subspace of some vector space. For instance, the set of all points on astraight line in  Rn passing through the origin is a subspace of  Rn. On the other hand, a straightline that does not pass through the origin is not a subspace of  Rn.

Let   W   be a set in a linear space   V . A vector   v   ∈   V   is said to be a   linear combination   of vectors in  W   if there exists a finite set of elements  w1, . . . ,  wn  ∈ W  and a finite set of scalars  α1,

. . . , αn ∈ F   such thatv  =  α1w1 + · · · + αnwn.

Let   W   be a nonempty subset of a linear space   V   , and denote by span(W ) the set of all linearcombinations of the vectors from  W . Then span(W ) is a linear subspace of  V  , called the   linear subspace generated by the set  W , or simply, the  span of  W ; the set  W   is said to   span   this linearsubspace. For example, the set {(1, 1, 0), (0, 0, 1)} ⊂  R3 spans the plane  {(x1, x2, x3)  ∈  R3 :  x1  =x2}, which is a subspace of  R3. It is readily verified that span(W ) is the smallest linear subspaceof a vector space  V   containing the subset W   of  V  .

Page 30: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 30/119

EE 580 — Linear Control Systems    3

4.3 Bases for Vector Spaces [1, Appendix A.2]

4.3.1 Linear Independence

Definition 4.4   Let  W   = {v1, . . . , vm}  be a finite nonempty set in a vector space  V . If there exist scalars  α1, . . . ,  αm, not all zero, such that 

α1v1 + · · · + αmvm = 0,

then the set  W  is said to be   linearly dependent over  F . If a set is not linearly dependent, then it is said to be   linearly independent. An infinite set of vectors  W   in  V   is said to be linearly independent if every finite subset of  W   is linearly independent.

If  U  is a finite set in a linear space  V   , then it is easily shown that  U  is linearly independent if and only if there is no proper subset  Z  of  U   such that span(Z ) = span(U ). For example, while theset {(1, 0), (0, 1), (2, −1)} ⊂ R2 is linearly dependent over the real field as 2(1, 0) − (0, 1) − (2, −1) =0, the sets   {(1, 0), (0, 1)},   {(1, 0), (2, −1)}, and   {(0, 1), (2, −1)}   are all linearly independent; inparticular, span{(1, 0), (2, −1)}  = span{(1, 0), (0, 1), (2, −1)}  = R2.

4.3.2 Bases

Definition 4.5   A set  W   in a linear space  V   is called a   basis   for  V   if  W   is linearly independent and if  span(W ) = V .

For example, each of the sets   {(1, 0), (0, 1)},   {(1, 0), (2, −1)},   {(0, 1), (2, −1)}, . . ., is a basisfor  R2, but none of the sets  {(1, 0), (−2, 0)},  {(1, −1), (−2, 2)}, . . ., is. Let  B  = {v1, . . . , vn}  be abasis for a linear space  V . Then it is easily shown that for each vector  v  ∈  V  , there exist uniquescalars α1, . . . , αn such that v  =  α1v1+· · ·+αnvn (Proof . If  v  =  α1v1+· · ·+αnvn  =  β 1v1+· · ·+β nvn,then (α1 − β 1)v1 + · · · + (αn − β n)vn  = 0, so by the linear independence of  B  we have  α1 = β 1, . . . ,αn  =  β n.); these scalars are called the  coordinates   of  v   with respect to the basis  B. Moreover, if {u1, . . . , um}  is any linearly independent set of vectors in  V  , then  m ≤ n; each basis of  V   consists

of exactly  n elements. If the number of elements in a basis for a linear space V   is  n < ∞, then  V is said to be of  n-dimensional , or finite-dimensional , and we write dim V   = n. Otherwise, V   is saidto be  infinite-dimensional .

Suppose either  F   =  R  or  F   =  C. Consider the n-dimensional vector space  F n over  F   and letB = {v1, . . . , vn} be a basis for F n. Then associated with every x ∈ F n is a unique n-tuple of scalars(α1, . . . , αn) relative to the basis  B  such that  x  =  α1x1 + · · · + αnxn. The vector  α = (α1, . . . , αn)is called the  coordinate representation  of the vector  x with respect to the basis  B . In particular, if 

e1 =

10...0

,   e2 =

01...0

, . . . ,   en  =

00...1

are the columns of the identity matrix   I  ∈   F n×n, then the set  {e1, . . . , en}   is called the   natural basis   (or the  standard basis ) for  F n; if  x  = (x1, . . . , xn)  ∈  F n, then the vector   x  = [x1   . . . xn]T

(i.e.,  x itself) is the coordinate representation of  x  with respect to the standard basis.As another example, consider the vector space of all polynomials   p :  C   →   C   of degree less

than   n  with real coefficients over the field of real numbers. A basis for this space is then givenby   B   =   {1, s , . . . , sn−1}, where   s   is a complex variable. Associated with each element   p(s) =α0 +  α1s + · · · + αn−1sn−1 is the unique  n-vector  α  = (α0, . . . , αn−1) ∈  Rn, which constitutes thecoordinate representation of  p(s) with respect to the basis  B .

Page 31: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 31/119

EE 580 — Linear Control Systems    4

4.4 Linear Transformations [1, Appendix A.3]

4.4.1 Matrix Representations

Definition 4.6   Let  V   and  W  be vector spaces over the same field  F . A function (or map)  H :  V   →W   is said to be a   linear transformation  (or   linear map) if 

H (αx + βy) = αH (x) + βH (y)

 for all  α,  β  ∈ F  and for all  x,  y ∈ V  .

Let  V   and  W  be two vector spaces over the same field  F . Let  L(V, W ) be the set of all lineartransformations from V   into  W . Define the sum of  S , T  ∈ L(V, W ) by

(S  + T )(v) = S (v) + T (v)

for all  v  ∈ V . Also, define the multiplication of  T  ∈ L(V, W ) by a scalar  α ∈ F   by

(αT )(v) = αT (v)

for all  v  ∈  V . Then  L(V, W ) is a vector space over  F , called the   space of linear transformations .For example, a function   f :  R2 →   R   is linear if and only if there exist   a1,   a2   ∈   R   such thatf (x1, x2) =   a1x1  +  a2x2   for all (x1, x2)   ∈   R2, and the set of all real-valued functions of realvariables of the form  f (x1, x2) =  a1x1 +  a2x2   (i.e., the set of all 1-by-2 real matrices [a1  a2]) is avector space; obviously, this vector space is nothing but  R1×2.

Theorem 4.7   Let  (V, F )  and  (W, F )  denote  n-dimensional and  m-dimensional vector spaces, re-spectively; let  BV    = {v1, . . . , vn}  and  BW   = {w1, . . . , wm}  be bases for  V   and  W , respectively. If T  :  V   → W   is a linear transformation, then there exists a unique matrix representation  A ∈ F m×n

of  T  with respect to the bases  BV    and  BW   such that, for any  v ∈ V ,

v = v1   · · ·   vnα   and    T (v) = w1   · · ·   wm

β   (4.2)

implies β =  Aα   with    A =

a1   · · ·   an

,

where the columns  a1, . . . ,  an ∈ F m of  A  satisfy 

T (v1) =

w1   · · ·   wm

a1, . . . , T  (vn) =

w1   · · ·   wm

an.

Proof . Suppose (4.2) holds with  α = [α1   · · ·   αn]T. Then, since  T  is linear, we have

T (v) = T 

v1   · · ·   vn

α

 =  T 

n

1αivi

 =

n

1αiT (vi) =

T (v1)   · · ·   T (vn)

α.

This, along with  T (v) =

w1   · · ·   wm

β, implies that  β =  Aα, where

A =

w1   · · ·   wm

−1 T (v1)   · · ·   T (vn)

.

Now the desired result follows from

w1   · · ·   wm

a1   · · ·   an

 =

T (v1)   · · ·   T (vn)

.  

A special case of Theorem 4.7 is as follows. With F  being a field, let  T   = (T 1, . . . , T  m) :  F n →F m be a linear transformation; let A ∈ F m×n be the unique matrix representation of  T  with respect

Page 32: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 32/119

EE 580 — Linear Control Systems    5

to the standard bases for  F n and  F m. Then, for all  x  = (x1, . . . , xn)  ∈  F n, we have  T (x) =  Ax

whenever  x = [x1   · · ·   xn]T is the column vector representation of  x; that is,

T (x) =

T 1(x1, . . . , xn)T 2(x1, . . . , xn)

...T m(x1, . . . , xn)

 =

T 1(1, 0, . . . , 0)   T 1(0, 1, . . . , 0)   · · ·   T 1(0, 0, . . . , 1)T 2(1, 0, . . . , 0)   T 2(0, 1, . . . , 0)   · · ·   T 2(0, 0, . . . , 1)

...  ...

  ...T m(1, 0, . . . , 0)   T m(0, 1, . . . , 0)   · · ·   T m(0, 0, . . . , 1)

   

A

x1

x2

...xn

   

x

.

For example, if  f  :  R2 → R2 is linear with f (1, 0) =

  3−1

 and f (0, 1) = [ 1

0], then we have f (x1, x2) =

  3−1

x1 + [ 1

0] x2   for all (x1, x2)  ∈  R2. Thus, this linear map  f  has matrix representation

  3 1−1 0

with respect to the standard bases for the domain  R2 and the codomain  R2. On the other hand, if we use the basis  B1  = {(1, −1), (1, 0)}   for the domain and the basis  B2  = {(1, 0), (0, −1)}   for the

codomain, then the linear map  f  has matrix representation

1 00  −1

−1 

  3 1−1 0

  1 1−1 0

 = [ 2 3

1 1].

4.4.2 Fundamental Theorem of Linear Equations

If  T  :  V   → W  is a linear transformation, the  null space   (or  kernel ) of  T   is defined as

 N (T ) = {v ∈ V  :  T (v) = 0},

and the range   (or  image ) of  T  is defined to be the set

R(T ) = T (V ) = {w ∈ W  :  w  =  T (v), v ∈ V }.

It is readily seen that   T   is injective (i.e., one-to-one) if and only if  N (T ) =  {0}, and that   T   issurjective (i.e., onto) if and only if  R(T ) = W .

Theorem 4.8   If  T  :  V   → W  is a linear transformation and if  V   is finite-dimensional, then  N (T )and  R(T )  are linear subspaces of  V   and  W , respectively, and satisfy 

dim N (T ) + dim R(T ) = dim V.

This result is called the fundamental theorem of linear equations  [2, Theorem 3.4]; the dimensionof  R(T ) is called the   rank   of  T   and denoted by rank T , and the dimension of  N (T ) is called thenullity   of   T . If   F   is a field and a linear map   T  :  F n →   F m has a matrix representation   A   ∈F m×n (with respect to the standard bases), so that  T (x) =  Ax   whenever  x   is the column vectorrepresentation of  x ∈ F n, then the range  R(A) and null space  N (A) of  A  are defined to be equalto R(T ) and N (T ), respectively. Denoted by rank A is the rank of  A, which is defined to be equalto rank T . The number of linearly independent columns (resp. rows) in  A  is called the column rank (resp.   row rank ) of  A; the column rank and the row rank of  A both coincide with the rank of  A.In fact, the span of the columns in   A   is equal to  R(A). For example,   A  =

  1   −2 1

−1 2 1

 has rank 2

(i.e., two linearly independent rows and two linearly independent columns) and so its null space

 N (A) = {(x1, x2, x3) ∈ C3 : x1 − 2x2 + x3 = 0,  −x1 + 2x2 + x3 = 0}

= span{(2, 1, 0)}

is of dimension 1, and its range

R(A) = {(y1, y2) ∈ C2 :  y1 =  x1 − 2x2 + x3, y2 = −x1 + 2x2 + x3, (x1, x2, x3) ∈ C

3}

= span{(1, −1), (1, 1)} = C2

has dimension 2.

Page 33: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 33/119

EE 580 — Linear Control Systems    6

4.4.3 Direct Sums of Linear Subspaces

Let  V  be a vector space and let  W   and U  be subsets of  V . The  sum  of  W   and  U   is defined by

W  + U  = {w + u :  w ∈ W, u ∈ U }.

In particular, if  W   and  U  are linear subspaces of  V , then  W  + U   is also a linear subspace. If  W and U  are linear subspaces of  V   such that U ∩ W   = {0}, then the sum W  + U  is called a  direct sum and denoted by W  ⊕ U . It is easily verified that  U  ∩ W   = {0}  if and only if, for every  v ∈ U  + W ,there exist unique elements  w  ∈ W   and  u ∈  U   such that  v  = u  + w. Let  BU   =  {u1, . . . , um}  andBW   = {w1, . . . , wl}  be bases for linear subspaces  U   and  W , respectively. Then  V   = U  ⊕ W   if andonly if  BU  ∪ BW  is a basis for  V . Consequently, dim(U  ⊕ W ) = dim U  + dim W .

If   x1,   x2  ∈  Rn are such that  xT

1x2  = 0, then we say   x1   and   x2   are   orthogonal   to each other.

It is easy to see that orthogonal vectors are linearly independent. Suppose W   and   U   are linearsubspaces of  Rn. If we have   wTu  = 0 for all   w  ∈  W  and for all   u  ∈  U , then  W   and  U   are saidto b e orthogonal to each other. In particular, if  W  ⊕ U   =  Rn and if  W   and  U  are orthogonal toeach other, then we write  W ⊥ = U   (and say  U   is the  orthogonal complement   of  W ) and U ⊥ = W .For example, if  W   = span{(1, 0)}  and  U  span{(1, 1)}, then we have that  W  + U   =  W  ⊕ U   =  R2;

however, W   and  U  are not orthogonal to each other because [ 10

]T [ 11

] = 1 = 0. On the other hand,if  W  = span{(−1, 1)} and  U  = span{(1, 1)}, then we have not only W  ⊕ U  = Rn but also W ⊥ = U .

Theorem 4.9   If  A ∈ Rm×n, then the following hold:

(a)   R(A) ⊕ N (AT) = Rm and  R(A)⊥ =  N (AT);

(b)   R(AT) ⊕ N (A) = Rn and  R(AT)⊥ =  N (A);

(c)   N (A) =  N (ATA);

(d)   R(A) = R(AAT).

Proof . To prove (a), Choose  y1  ∈ R(A) and  y2  ∈ N (AT). Then we have  y1   =  Ax   for somex ∈ Rn, so  yT

1y2  = (Ax)Ty2 = xT(ATy2) = 0. This implies  R(A) is orthogonal to  N (AT). Now,

choose a nonzero   y  ∈ N (AT). Then  ATy   = 0 and so   yTAx  = 0 for all   x  ∈  Rn. This implies

y = Ax  for any x ∈ Rn; that is, y  /∈ R(A). Hence R(A) ∩ N (AT) = {0}. Since rank A = rank AT,

Theorem 4.8 applied to   AT implies dim R(A) + dim N (AT) = dim R(AT) + dim N (AT) =   m.This means that, if   {v1, . . . , vk}   is a basis for   R(A) and if   {w1, . . . , wl}   is a basis for   N (AT),then   k  + l   =   m. Since   vi   and   w j   are orthogonal to each other for all   i   and   j, we have that{v1, . . . , vk, w1, . . . , wl} is a linearly independent set of vectors. Since R(A)⊕N (AT) is a subspaceof  Rm, this is possible only if  R(A) ⊕ N (AT) = R

m. Moreover, as  R(A) is orthogonal to  N (AT),we conclude that  R(A)⊥ = N (AT). This complete the proof of (a).

Replacing   A   in (a) with   AT yields (b) as well. To prove (c), suppose ATAx   = 0. Then wehave   xTATAx  =  Ax2 = 0, and so   Ax  = 0. Thus N (ATA)  ⊂ N (A). Conversely, if   Ax  = 0,then ATAx = 0 as well, so  N (A) ⊂ N (ATA). This proves (c). Lastly, it follows from (a)–(c) thatR(A) =  N (AT)⊥ =  N (AAT)⊥ = R(AAT), so that (d) holds true.  

By part (d) of Theorem 4.9, we have rank A  = rank AAT, and so   A   has full row rank (i.e.,rank A =  m) if and only if   AAT is nonsingular . Similarly, we have rank AT = rank ATA, so  A

has full column rank (i.e.,  rank A =  n) if and only if  ATA ∈ Rn×n is nonsingular .

Page 34: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 34/119

EE 580 — Linear Control Systems    7

4.4.4 Linear Algebraic Equations

Consider the system of linear algebraic equations of the form

Ax =  y,   (4.3)

where   A  ∈ Rm×n

and   y  ∈ Rm

are given and   x  ∈ Rn

is to be determined. For a given   y, thereexists at least one solution   x   to (4.3) if and only if   y  ∈ R(A). Every solution   x   to (4.3) can beexpressed as a sum x  =  x p + xh, where x p  is a particular solution of (4.3) and  xh  satisfies Axh  = 0.Thus, the solution to (4.3) is unique if and only if  N (A) = {0}. The following are consequences of Theorem 4.9, which are presented without proof.

Theorem 4.10   (Least-Squares Solution)  A matrix   A ∈ Rm×n has full row rank (i.e.,  m ≤ n  and 

rank A =  m) if and only if it represents a surjective linear map. In this case,  AAT is nonsingular,and   x0 =  AT(AAT)−1y   is such that  Ax0 =  y  and  x0 ≤ x  for all solutions   x  to   (4.3).

Theorem 4.11   (Least-Squares Approximation)   A matrix   A  ∈  Rm×n has full column rank (i.e.,m ≥ n  and  rank A =  n) if and only if it represents an injective linear map. In this case,  ATA   is 

nonsingular, and  x0 = (ATA)−1ATy   is such that  y − Ax0 ≤ y − Ax  for all  x ∈ Rn.

Theorem 4.12   (Unique Solution)  A matrix  A ∈ Rn×n has full row and column ranks (i.e.,  m  =  n

and  rank A =  n) if and only if  A   represents a bijective linear map. In this case,  A  is nonsingular,and   x0 =  A−1y   is the unique solution to   (4.3).

In Theorem 4.10, because  x0 =  AT(AAT)−1y is such that  x0 − 0 ∈ R(AT) =  N (A)⊥, we havexT(x0 − 0) = 0 for all  x ∈ N (A). For this reason, we say that the origin is the  orthogonal projection of  x0  onto  N (A). On the other hand, in Theorem 4.11, b ecause  x0  = (ATA)−1ATy   is such thatAx0 − y ∈ N (AT) = R(A)⊥, we see that  zT(Ax0 − y) = 0 for all  z ∈ R(A). For this reason, wesay that  Ax0  is the orthogonal projection of  y  onto  R(A).

4.5 Equivalence and Similarity [1, Appendix A.4]

4.5.1 Change of Bases: Vector Case

Let A ∈ F m×n be the matrix representation of a linear transformation that maps an  n-dimensionalvector space (V, F ) into an  m-dimensional vector space (W, F ). Let  {v1, . . . , vn}  be a basis for  V and let  {v1, . . . , vn}  be a set of vectors in  V   given by

v1   · · ·   vn

 =

v1   · · ·   vn

P   (4.4a)

with

P = p1   · · ·   pn =

 p11   · · ·   p1n

...   . . .   ... pn1   · · ·   pnn

.   (4.4b)

The set  {v1, . . . , vn} forms a basis for  V  if and only if the matrix  P  = ( pij) ∈ F n×n is nonsingular .(Proof . If   P   is nonsingular, then [v1   · · ·   vn] is nonsingular, and so sufficiency follows. If   pn   =α1p1 +· · ·+αn−1pn−1, then direct computation shows vn  = [v1   · · ·   vn]pn  = α1v1 +· · ·+αn−1vn−1,so  {v1, . . . , vn}   cannot b e linearly independent. This proves necessity.) We call   P   the  matrix of the basis   {v1, . . . , vn}   with respect to the basis   {v1, . . . , vn}. Here, the   jth column   p j   of   P   isthe coordinate representation of  v j   with respect to the basis  {v1, . . . , vn}. Given a   v   ∈   V , let

Page 35: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 35/119

EE 580 — Linear Control Systems    8

B  =  {v1, . . . , vn}  and  B  =  {v1, . . . , vn}  be bases for the vector space  V . Suppose that α  and  αare the coordinate representations of  v  with respect to the bases  B  and  B, respectively; that is,

v =

v1   · · ·   vn

α =

v1   · · ·   vn

 α.

Then it follows from  P = [v1   · · ·   vn]−1[v1   · · ·   vn] that we have

Pα =  α,

where  P  is the matrix of the basis  B  with respect to the basis  B   satisfying (4.4).

4.5.2 Change of Bases: Matrix Case

Let   V   and   W   be   n-dimensional and   m-dimensional vector spaces, respectively, over a commonfield  F , and let  {v1, . . . , vn}  and  {w1, . . . , wm} be bases for  V   and  W , respectively. Suppose that{v1, . . . , vn}   is another basis for  V   with  P ∈  F n×n being the matrix of  {v1, . . . , vn}  with respectto  {v1, . . . , vn}, and that  {w1, . . . , wm}   is another basis for  W   with  Q ∈ F m×m being the matrixof  {w1, . . . , wm} with respect to  {w1, . . . , wm}. Then we have

v1   · · ·   vn

 =

v1   · · ·   vn

P,

w1   · · ·   wm

Q =

w1   · · ·   wm

.

Let  T  :  V   →  W  be a linear transformation. If  A  ∈  F m×n is the matrix of  T  with respect to thebases   {v1, . . . , vn}   and  {w1, . . . , wm}, and if  A   ∈  F m×n is the matrix of   T   with respect to thebases  {v1, . . . , vn}  and  {w1, . . . , wm}, then it follows from   y  =  T (x) with   y  = [w1   · · ·   wm]β  =[w1   · · ·  wm]β,  x = [v1   · · ·   vn]α = [v1   · · ·   vn]α, and  β =  Aα  that we have  β =  Aα  with

A =  QAP.   (4.5)

Definition 4.13   Let  F  be a field. A matrix  A ∈ F m×n is said to be   equivalent   to an   A ∈  F m×n

if there exist nonsingular matrices  Q ∈ F m×m and  P ∈ F n×n such that   (4.5)  is satisfied.

It can be shown that two matrices   A  and  B  are equivalent if and only if rank A  = rank B; thatis,   if   A,   B  ∈  Rm×n, then   rank A  = rank B   if and only if there exist nonsingular   P  ∈  Rn×n and Q ∈ Rm×m such that  B =  QAP.

4.5.3 Similarity Transformation

Let  V   be an  n-dimensional vector space over a field  F , and let  {v1, . . . , vn}  and  {v1, . . . , vn}  bebases for V . Let P ∈ F n×n be the matrix of {v1, . . . , vn} with respect to {v1, . . . , vn}. If  T  :  V   → V is a linear transformation, and if  A ∈  F n×n is the matrix of  T   with respect to  {v1, . . . , vn}, then

the matrix  A ∈ F n×n

of  T  with respect to  {v1, . . . , vn}  is given by

A =  P−1AP.   (4.6)

Definition 4.14   Let  F   be a field. A matrix  A  ∈  F n×n is said to be   similar   to an   A  ∈  F n×n if there exists a nonsingular matrix   P ∈ F n×n such that   (4.6)   is satisfied. The function  f :  F n×n →F n×n defined by   f (A) =   P−1AP   for some nonsingular   P   and for all   A   is called a   similaritytransformation.

Page 36: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 36/119

EE 580 — Linear Control Systems    9

4.5.4 Polynomials of Linear Transformations

If  T  :  Cn →  Cn is a linear transformation, then we write  T 0 =  I  (i.e., the identity map such thatIx =  x  for all  x ∈ Cn),  T 1 = T ,  T 2 = T T   = T (T (·)), and so on. In general, if  p(s) is a polynomialof degree  m   in  s  given by

 p(s) = α0 + α1s + · · · + αmsm,   (4.7)

then we write  p(T ) = α0I  + α1T 1 + · · · + αmT m. It is readily seen that, whenever  A :  Cn×n is thematrix of  T  with respect to the standard basis, then  p(A) defined by

 p(A) = α0I + α1A + · · · + αmAm

is the matrix of  p(T ) with respect to the standard basis. For   A,   P  ∈  Cn×n, if   P   is nonsingular

and  p  is as in (4.7), we have p(P−1AP) = P−1 p(A)P;

that is,   whenever   B   is similar to   A  under a similarity transformation,   p(B)   is similar to   p(A)under the same similarity transformation .

4.5.5 Change of Bases for State-Space Models

Consider an LTI system of the form

x(t) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t),(4.8)

where, for each   t ∈ R, the column vectors  x(t) ∈ Rn,  u(t) ∈ Rm, and  y(t) ∈ Rl are the coordinaterepresentations of the state, input, and output of the system with respect to the standard bases forthe state space Rn, input space Rm, and output space Rl, respectively. If  x(t),  u(t), and y(t) are thecoordinate representations of the state, input, and output with respect to new bases  {v1, . . . , vn},{w1, . . . , wm}, and {z1, . . . , zl}, respectively, then we have

Px(t) = x(t),   Qu(t) = u(t),   and   Ry(t) = y(t)

for  t ∈ R, where

P =

v1   · · ·   vn

,   Q =

w1   · · ·   wm

,   and   R =

z1   · · ·   zl

.

Thus, in terms of the new bases, the state-space model (4.8) is represented as

˙x(t) = (P−1AP)x(t) + (P−1BQ)u(t),

y(t) = (R−1CP)x(t) + (R−1DQ)u(t).(4.9)

If the transfer function matrix of the original model (4.8) is H(s) = C(sI − A)−1B + D, then thetransfer function matrix of the new model (4.9) is given by   R−1 H(s)Q. As we have seen earlier,the two models are equivalent if  Q =  I  ∈  Rm×m and  R  =  I  ∈  Rl×l. In general, since there exists

one-to-one correspondence between the two models, one can perform analysis and synthesis for (4.9)and transform the results back to the original spaces to obtain the corresponding results for (4.8).

References

[1] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[2] S. Axler, Linear Algebra Done Right , 2nd ed. New York, NY: Springer, 1997.

Page 37: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 37/119

EE 580 — Linear Control Systems

V. Structural Properties of Matrices

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

5.1 Definitions and Facts

5.1.1 Determinants

Given a square matrix  A ∈ Cn×n, the determinant of  A, denoted by det A, can be defined induc-

tively in the following way. Define the determinant of a 1-by-1 matrix to be the value of its singleentry. Assume the determinant is defined over  C(n−1)×(n−1) and let  Aij  ∈ C

(n−1)×(n−1) denote thesubmatrix of  A = (aij) ∈ C

n×n resulting from the deletion of row  i  and column  j. Then

n j=1

(−1)i+ jaij det Aij  =ni=1

(−1)i+ jaij det Aij

for all   i,   j   ∈   n, and this common value is defined to be the determinant of   A. It is clear thatdet AT = det A by definition. Also, it is also easily verified that det I = 1 and det 0 = 0, and thatdet A = 0 if and only if  A   is nonsingular. The following are useful facts [1, A.11 & 12]:

•   If  A,  B ∈ Cn×n, then det AB = det A det B. (This does not hold for nonsquare  A and  B.)

•   If  A  and  B  are square and if  A   is nonsingular,

det

A B

C D

 = det A det(D − CA−1B).

(Recall that   D − CA−1B   is the Schur complement of   A.) In particular, if   B   =   0, then(regardless of the invertibility of  A) we have

det

A 0

C D

 = det A det D.

•   If  A ∈ Cm×n and  B ∈ Cn×m, then det(I − AB) = det(I − BA).

5.1.2 Traces

If  A ∈ Cn×n, then the trace of  A, denoted by tr A, is the sum of the diagonal entries of  A; that is,

tr A =n

i=1 aii   for  A = (aij). Clearly, trace is a linear operation: tr(αA + β B) = α tr A + β  tr B

for all  α,  β  ∈  C  and for all   A,   B  ∈  Cn×n. If  A  ∈  C

m×n and   B  ∈  Cn×m, then direct calculation

reveals that tr AB = tr BA.

1

Page 38: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 38/119

EE 580 — Linear Control Systems    2

5.1.3 Eigenvalues and Eigenvectors

Let (Cn,C) be the n-dimensional complex vector space (over the complex field). Let A ∈ Cn×n be a

matrix representation of a linear transformation that maps Cn into  Cn. If a complex number  λ ∈ C

and a nonzero complex vector  v ∈ Cn are such that  Av =  λv, then  λ   is called an  eigenvalue  of  A

and   v   is called an  eigenvector   of   A  associated with the eigenvalue   λ. The   fundamental theorem 

of linear equations   [2, Theorem 3.4] says that the sum of the rank and nullity of   A   equals   n.Consequently, the linear equation Av  =  λv, or equivalently (λI−A)v = 0, has a nonzero solution  v

if and only if rank(λI−A) < n (i.e., λI−A is singular). Thus, the eigenvalues of  A  are the complexnumbers s  that satisfy

det(sI − A) = 0.   (5.1)

We call det(sI − A), which is a polynomial of degree  n   in  s, the  characteristic polynomial   of   A,and (5.1) the  characteristic equation   of  A. The well-known, but nontrivial,  fundamental theorem of algebra   [2, Theorem 4.7] implies that a polynomial of degree   n   with complex coefficients hasexactly n roots, counting multiplicities, among the complex numbers. A consequence of this theoremis that an   n-by-n   matrix has exactly   n   eigenvalues, counting (algebraic) multiplicities. If   λ1,. . .   λn ∈ C  are these eigenvalues, then the following hold true [3, Theorem 1.2.12]:

tr A =ni=1

λi   and det A =ni=1

λi.

If  λ1, . . . ,  λ p   are distinct eigenvalues of   A  ∈  Cn×n (where  p  ≤  n), then the dimension of the

linear subspace  N (λiI − A) in  Cn (i.e., the number of linearly independent eigenvectors associatedwith the eigenvalue λi over the complex field) is called the geometric multiplicity  of the eigenvalue λi.On the other hand, if the characteristic polynomial is given by det(sI − A) =

  pi=1(s − λi)

mi , where pi=1 mi   =  n, then   mi   is the   algebraic multiplicity  of the eigenvalue   λi. If  f (s) is a polynomial

in  s, then the roots of the characteristic polynomial of  f (A) are  f (λ1), . . . ,  f (λ p), and det(sI −f (A)) =

  pi=1(s − f (λi))mi ; that is, each  f (λi) is an eigenvalue of  f (A) of algebraic multiplicity

mi, i  = 1, . . . ,  p. Also, since det(sI− A) = det P

−1

(sI− A)P = det(sI− P

−1

AP) for any invertibleP ∈  Cn×n, similar matrices have the same characteristic polynomial; that is, they have the same

set of eigenvalues, counting (algebraic) multiplicities. Finally, the eigenvalues of  AT are the sameas those of  A, counting multiplicities, because det(λI − AT) = det(λI − A)T = det(λI − A).

5.1.4 An Example

To illustrate the definitions and facts presented so far, consider

A =

1 0 1 00 0 0 0

−1 0 1 00 1 0 0

,   P =

0 0 0 10 1 0 01 0 0 00 0 1 0

=

0 0 0 10 1 0 01 0 0 00 0 1 0

−1

.

If  {e1, e2, e3, e4}   is the standard basis for  C4, then we have [e4  e2  e1  e3] = [e1  e2  e3  e4]P =  P, sothe change of basis defined by  P  is nothing but a reordering of the basis vectors. In particular, AP

amounts to a reordering of the columns of   A  and   P−1A   corresponds to a reordering of the rowsof  A. We have

P−1AP =

0 1 0 00 0 0 00 0 1 10 0   −1 1

=

A1   0

0 A2

  with   A1 =

0 10 0

,   A2  =

 1 1−1 1

,   0 =

0 00 0

,

Page 39: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 39/119

EE 580 — Linear Control Systems    3

so it follows from det(sI − A) = det(sI − P−1AP) = det(sI − A1)det(sI − A2) that the set of eigenvalues of  A is the union of the sets of eigenvalues of  A1  and  A2. Since det(sI − A1) = s2, wehave that 0 and 0 are eigenvalues of  A1. Also, since det(sI−A2) = (s−1)2 +1, we have that 1+i and1 − i are eigenvalues of  A2. Thus, 0, 1+ i, and 1 − i are the distinct eigenvalues of  A; their algebraicmultiplicities are 2, 1, and 1, respectively. While the geometric multiplicities of eigenvalues 1 + i

and 1 − i   are both equal to 1, it follows from  N (0I − A) =  {x   ∈  C

4

:  x1  +  x3   = 0, x1 − x3   =0, x2   = 0}  =  {x  ∈  C4 :  x1   =  x2   =  x3   = 0}  = span{(0, 0, 0, 1)}  that the geometric multiplicity of 

eigenvalue 0 (i.e., the dimension of the subspace spanned by the eigenvectors of  A  corresponding toeigenvalue 0) is 1, which is less than its algebraic multiplicity. Finally, counting multiplicities, theset of all eigenvalues of  A  is {0, 0, 1 + i, 1 − i}, so indeed tr A =  a11 + a22 + a33 + a44 = 1+0+1+ 0 =0 + 0 + (1 + i) + (1 − i) and det A = det A1 det A2 = 0 · 2 = 0 · 0 · (1 + i) · (1 − i).

5.2 Upper-Triangular Form of Matrices [2, Chapter 5]

5.2.1 Invariant Subspaces

Definition 5.1   Let   V    be a vector space; let   T  :  V   →   V   be a linear transformation. A linear subspace  U   of  V   is said to be   invariant  under  T   if  u ∈ U   implies  T u ∈ U .

If   A   ∈   Cn×n is a matrix representation of a linear map   T  :  Cn →   C

n, and if a subspace   U of   Cn is invariant under   T , then we say   U   is invariant under   A. Clearly, the spaces   {0},   Cn,

 N (A), and  R(A) are all invariant under  A. In general, if  p(s) is a polynomial in  s, then it followsimmediately from p(A)A =  A p(A) that  N 

 p(A)

 and R

 p(A)

 are both invariant under  A.

5.2.2 Upper-Triangular Form

Suppose A ∈ Cn×n and  {v1, . . . , vn}  is a basis of  Cn. Recall that  P = [v1   · · ·   vn] is the matrix of 

the basis {v1, . . . , vn} with respect to the standard basis  {e1, . . . , en} such that the representationof  A  with respect to  {v1, . . . , vn}  is given by the similarity transform  A  = (aij) =  P−1AP. The

matrix  ¯A is called an upper-triangular matrix representation  of  A if all the entries below its diagonalare equal to zero (i.e. aij  = 0 for  i > j ).

Lemma 5.2   Let  A ∈ Cn×n; let  {v1, . . . , vn}  be a basis of  Cn. The following are equivalent:

(a)   P−1AP  is upper triangular with  P = [v1   · · ·   vn];

(b)   Avk  ∈ span{v1, . . . , vk}   for each  k  = 1, . . . ,  n;

(c)   span{v1, . . . , vk}  is invariant under  A   for each  k = 1, . . . ,  n.

Proof . We will show (a) ⇒ (b) ⇒ (c) ⇒ (a). Condition (a) is equivalent to the existence of a11,a12, a22, . . . , a1n, . . . , ann ∈ C such that

A

v1   · · ·   vn

 =

v1   · · ·   vn

a11   a12   · · ·   a1n

0 a22   · · ·   a2n...

  ...  . . .

  ...0 0   · · ·   ann

.   (5.2)

Thus, if (a) holds, then   Avk   = a1kv1  + · · · + akkvk   ∈   span{v1, . . . , vk}   for all   k, and so (b)holds. If (b) holds, then we have   Av1  ∈  span{v1},   Av2  ∈  span{v1, v2}, and so on. This meansAv1, . . . ,   Avk  belong to span{v1, . . . , vk}   for each   k. Choose any   x  ∈   span{v1, . . . , vk}. Then

Page 40: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 40/119

EE 580 — Linear Control Systems    4

x   =   α1v1  + · · ·  +  αkvk   for some   α1, . . . ,   αk   ∈   C, and hence we have   Ax   =   α1Av1  + · · · +αkAvk  ∈  span{v1, . . . , vk}. This implies (c) holds true. Finally, if (c) holds, then   Av1  = a11v1,Av2 = a12v1 + a22v2, and so on for some scalars aij . This gives (5.2), so (a) holds.  

Lemma 5.3   Let   A ∈  Cn×n. Let   V  ∈  Cn×k have full column rank. Then the columns of   V   span an invariant subspace of  Cn under  A   if and only if  AV =  VAV    for some   AV   ∈ Ck×k. Moreover,all eigenvalues of  AV   are eigenvalues of  A.

Proof . We need to show necessity (i.e., the “only if” part) and sufficiency (i.e., the “if” part)separately. To show necessity, suppose  V  = [v1   · · · vk] has linearly independent columns that spanan invariant subspace of  Cn under A. Then for each  j  = 1, . . . , k  there exist a1 j, . . . , akj  ∈ C suchthat  Av j  = a1 jv1 + · · · + akjvk. Thus letting  AV  be the matrix whose entry (i, j) is aij  yields thatAV  =   VAV  . Conversely, to show sufficiency, suppose  AV  =  VAV  . Then, for each   u  ∈  C

k, thevector  Vu   is a linear combination of  v1, . . . ,  vk. Thus it follows from  A(Vu) =  V(AV u) =  Vw,with  w =  AV u ∈ Ck, that the image of each linear combination of  v1, . . . ,  vk  under  A  is anotherlinear combination of  v1, . . . ,  vk, and hence that span{v1, . . . , vk}   is invariant under  A. Finally,if AV v =  λv for some λ ∈ C and some nonzero v ∈ Ck, then A(Vv) = V(AV v) = V(λv) = λ(Vv),

so  λ   is an eigenvalue of  A with  Vv being a corresponding eigenvector of  A.  

Theorem 5.4  Every square matrix   A  ∈  Cn×n has an upper-triangular matrix representation  A.

Moreover, the eigenvalues of  A  consist precisely of the entries on the diagonal of  A.

Proof . We will prove the result by induction on  n. Clearly the desired result holds if  n  = 1.As an induction hypothesis, suppose that  n >  1 and that the result holds for all  A  ∈  C

1×1, . . . ,A ∈  C

(n−1)×(n−1). We will complete the proof by showing that, under this hypothesis, the resultholds for all  A ∈  C

n×n as well. Choose any  A ∈  Cn×n. Let  λ  be an eigenvalue of  A. Since there

exists a nonzero vector   v   (i.e., an eigenvector of   A  associated with   λ) in  N (λI − A), and sincedim R(λI − A)+dim N (λI − A) = n, we have that dim R(λI − A) < n. Moreover, R(λI − A) is aninvariant subspace of  Cn under A. By Lemma 5.3, there exists a basis  {u

1, . . . , u

k} of  R(λI − A),

with  k < n, such that putting  U  = [u1   · · · uk]  ∈  Cn×k yields  AU =  UAU   for some  AU   ∈  C

k×k.Then, by the induction hypothesis and by Lemma 5.2, there exists a basis {v1, . . . , vk} for  Ck suchthat   AU vi   ∈   span{v1, . . . , vi}   for each   i  = 1, . . . ,  k. Also, if we extend  {v1, . . . , vk}  to a basis{v1, . . . , vk, vk+1, . . . , vn} of  Cn, then for each  i =  k  + 1, . . . , n  we have  Avi  = λvi − (λI − A)vi,where (λI − A)vi  ∈  span{v1, . . . , vk}, and so  Avi  ∈  span{v1, . . . , vi}. Therefore, by Lemma 5.2,the matrix  A   has an upper-triangular representation  A =  P−1AP  with  P = [v1   · · · vn]  ∈  Cn×n.If the entries on the diagonal of  A are  λ1, . . . ,  λn, then the corresponding entries of  λI −  A equalλ − λ1, . . . , λ − λn. Since λI −  A =  P−1(λI − A)P (i.e., λI −  A and  λI − A are similar), and sincedet(λI −  A) =

 ni=1(λ − λi) (due to  A being upper triangular), the matrix  λI − A is not invertible

(i.e., λ  is an eigenvalue of  A) if and only if  λ  equals one of the  λ j’s. This completes the proof.  

This result is called the   Schur triangularization theorem . It will be used in Section 5.3.2 to

prove the Cayley-Hamilton Theorem, which is one of the most important linear algebraic tools.

5.2.3 Diagonal Form

Theorem 5.5   If   A  ∈  Cn×n has   p   distinct eigenvalues  λ1, . . . ,   λ p  ∈  C, and if   v1, . . . ,   v p  ∈  Cn

are corresponding eigenvectors, then  {v1, . . . , v p}   is linearly independent.

Proof . To prove the result by contradiction, suppose {v1, . . . , v p} is linearly dependent. Let k  bethe smallest positive integer satisfying vk  ∈ span{v1, . . . , vk−1}, so that vk  = α1v1 + · · ·+αk−1vk−1

Page 41: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 41/119

EE 580 — Linear Control Systems    5

for some  α1, . . . ,  αk−1 ∈ C. Then we have both

Avk  = α1Av1 + · · · + αk−1Avk−1 =  α1λ1v1 + · · · + αk−1λk−1vk−1

andAvk  = λkvk  = λkα1v1 + · · · + λkαk−1vk−1,

which give   α1(λk  − λ1)v1  + · · · + αk−1(λk  − λk−1)vk−1   = 0. By construction,   {v1, . . . , vk−1}   islinearly independent, and so we must have  α1   =   · · ·  =  αk−1   = 0. However, this implies   vk   = 0,which contradicts the fact that  vk   is an eigenvector of  A.  

This theorem suggests the following: If  A ∈ Cn×n has  n  distinct eigenvalues  λ1, . . . , λn, and if 

v1, . . . ,  vn ∈ Cn are corresponding eigenvectors of  A, then {v1, . . . , vn}  is a basis for  Cn. That is,

Cn = span{v1, . . . , vn} =  N (λ1I − A) ⊕ · · · ⊕ N  (λnI − A)

over the complex field. Moreover, if we let

P =

v1   · · ·   vn

,

then  P  is the matrix of the basis {v1, . . . , vn} with respect to the standard basis {e1, . . . , en} suchthat the representation of  A with respect to  {v1, . . . , vn} is given by

A =  P−1AP = diag{λ1, . . . , λn} =

λ1   0   · · ·   00   λ2   · · ·   0...

  ...  . . .

  ...0 0   · · ·   λn

.   (5.3)

This is because the  ith column of  AP  equals  Avi, which equals  λvi, which equals the  ith columnof  PA. The matrix  A   in (5.3) is called a  diagonal matrix representation   of  A, which is a specialtype of upper-triangular representations. The existence of  n  distinct eigenvalues is sufficient but

not necessary for an  n-by-n  matrix to be diagonalizable like this. Some of the exact conditions fordiagonalizability proved in [2, Theorem 5.21] are as follows:

Theorem 5.6   Let  λ1, . . . ,  λ p ∈ C  be the distinct eigenvalues of  A ∈ Cn×n. Each of the following 

is equivalent to the existence of a diagonal matrix representation of  A:

(a) The geometric multiplicity is the same as the algebraic multiplicity for each  λi,  i = 1, . . . ,  p;

(b)   Cn has a basis consisting of eigenvectors of  A;

(c)   Cn =  N (λ1I − A) ⊕ · · · ⊕ N  (λ pI − A).

5.2.4 An Example

Let  A,  P ∈ C4×4 and  A1,  A2 ∈ C

2×2 be as in Section 5.1.4. Then we have

AP =  P

A1   0

0 A2

  ⇒   A

e4   e2

 =

e4   e2

A1   and   A

e1   e3

 =

e1   e3

A2,

so Lemma 5.3 implies that both span{e4, e2}   and span{e1, e3}   are invariant subspaces under   A.Indeed, a direct computation also implies span{e2, e4}, for instance, is invariant under  A  becausex  =   αe2 +  β e4   = [0   α   0  β ]T gives   Ax  = [0 0 0   α]T =   αe4   ∈   span{e2, e4}. While   A1   is upper

Page 42: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 42/119

EE 580 — Linear Control Systems    6

triangular, A2 is not, so even though  P−1AP is block diagonal, it is not upper triangular. However,the two eigenvalues 1+i and 1−i of  A2 are distinct and hence by Theorem 5.5 the eigenvectors of  A2

span C2 and A2 is diagonalizable. More specifically, we have  A2v1  = (1+i)v1 and  A2v2 = (1−i)v2,

where v1 = [1 i]T and v2  = [1   − i]T are linearly independent, so  V−1A2V = diag{1 + i, 1 − i} withV = [v1  v2]. Therefore, as

Q =  P

I 0

0 V

 =

0 0 1 10 1 0 00 0   i   −i1 0 0 0

Q−1AQ =

I 0

0 V−1

A1   0

0 A2

I 0

0 V

=

A1   0

0 V−1A2V

 =

0 1 0 00 0 0 00 0 1 + i   00 0 0 1 − i

,

we have that  Q−1AQ  is upper triangular. Note that, according to Theorem 5.6, the matrix A   isnot diagonalizable because, while the eigenvalue 0 has an algebraic multiplicity of 2, its geometricmultiplicity is only 1.

5.3 Jordan Canonical Form of Matrices [2, Chapter 8]5.3.1 Generalized Eigenvectors

Due to Theorem 5.6, if  A ∈ Cn×n does not have enough number of linearly independent eigenvec-

tors, there is no diagonal matrix representation of  A. In such cases, we generalize the concept of eigenvectors in order to generate a basis  B   for  Cn such that the representation of  A  with respectto  B  is upper triangular.

Definition 5.7   Let  λ  be an eigenvalue of   A ∈ Cn×n. A nonzero vector   v ∈  Cn is called a  gener-alized eigenvector  of  A   corresponding to  λ  if  (λI − A)kv = 0   for some positive integer  k.

Every eigenvector is a generalized eigenvector, but the converse is not true. For example, the

matrix  A ∈ C4×

4 considered in Sections 5.1.4 and 5.2.4 is such that  A2[0 x2 0 x4]T = 0 for all  x2,x4 ∈ C. While (0, 1, 0, 0) and (0, 0, 0, 1) are both generalized eigenvectors of  A  corresponding to theeigenvalue 0, the former is not an eigenvector of  A because  Ax = 0 only if  x ∈ span{(0, 0, 0, 1)}.

It is easily seen that, for any  A ∈ Cn×n, we have

{0} =  N (A0) ⊂ N (A1) ⊂ · · · ⊂ N  (An) =  N (An+1) = · · ·   ,

Cn = R(A0) ⊃ R(A1) ⊃ · · · ⊃ R(An) = R(An+1) = · · ·   .

If  λ  is an eigenvalue of  A ∈ Cn×n, then replacing A  with λI − A we see that the span of generalized

eigenvectors of  A  associated with  λ  equals  N (λI − A)n. It is shown in [2, Theorem 8.10] that  the algebraic multiplicity of  λ  is precisely equal to the dimension of  N (λI − A)n. A matrix A ∈ C

n×n iscalled nilpotent   if  Ak = 0   for some positive integer  k. If  A  is nilpotent, then every nonzero vectorin Cn is a generalized eigenvector of  A  corresponding to the eigenvalue 0, and thus  N (An) = Cn. Itis shown in [2, Lemma 8.26] that every nilpotent matrix is similar to a “strictly” upper-triangularmatrix, whose entries on and below the diagonal are all 0’s.

5.3.2 Cayley-Hamilton Theorem

The following result, called the  Cayley-Hamilton Theorem , is very useful to us.

Theorem 5.8   If  p(s) = det(sI − A)  where  A ∈ Cn×n, then  p(A) = 0.

Page 43: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 43/119

EE 580 — Linear Control Systems    7

Proof . Suppose {v1, . . . , vn} is a basis for  Cn with respect to which the representation of  A  hasan upper-triangular form  A = (aij) as in (5.2). To prove  p(A) =  0, we need to show  p(A)vk  = 0for all   k   = 1, . . . ,   n. Since   p(A) is a constant multiple of 

 ni=1(λiI − A), where   λ1, . . . ,   λn

are the eigenvalues of   A, it suffices to show k

i=1(λiI − A)vk  = 0 for  k  = 1, . . . ,  n. To do thisby induction, let   k   = 1. Then we have   Av1   = a11v1, where a11   =   λ1   for some eigenvalue   λ1

of   A, giving (λ1I − A)v1  = 0. Now, as an induction hypothesis, suppose there exists a positiveinteger  j  such that k

i=1(λiI − A)vk  = 0 for  k  = 1, . . . ,  j . We will complete the proof by showing

that, under this hypothesis, we have  j+1

i=1 (λiI − A)v j+1   = 0 as well. Equation (5.2), togetherwith Theorem 5.4, gives (a j+1 j+1I − A)v j+1  ∈  span{v1, . . . , v j}, where a j+1 j+1  =  λ j+1   for some

eigenvalue λ j+1 of  A. Thus, by the induction hypothesis,  j

i=1(λiI−A) applied to (λ j+1I−A)v j+1

gives 0. This completes the proof.  

Corollary 5.9   If   A   ∈   F n×n with either   F   =   R   or   F   =   C, and if   f (s)   is a polynomial whose coefficients are in  F , then there exist  β 0, . . . ,  β n−1 ∈ F   such that 

f (A) = β 0I + β 1A + · · · + β n−1An−1.

Proof . If  p(s) = det(sI − A) is the characteristic polynomial of  A, then the coefficients of  p(s)are in F . Moreover, there exist two unique polynomials  q (s) and r(s), whose coefficients in  F , suchthat  f (s) =  p(s)q (s) +  r(s), where the degree of  r(s) is at most  n − 1. By the Cayley-HamiltonTheorem, we have  p(A) = 0, and so  f (A) = p(A)q (A) + r(A) = r(A).  

For example, let us use the Cayley-Hamilton theorem and its corollary to evaluate   At for all

t   = 0, 1, ..., where   A   =

1/2 10   −1/2

. The characteristic polynomial of   A   is given by   p(s) =

det(sI − A) = (s − 1/2)(s + 1/2). Let  f (s) = st. Then by Corollary 5.9 there are  β 0(t),  β 1(t) ∈ R

such that f (A) = β 0(t)I + β 1(t)A. Let r(s) = β 0(t) + β 1(t)s. Then  f (s) − r(s) has  p(s) as a factorif and only if  f (s) − r(s) = 0 for   s   = 1/2 and for   s  =  −1/2. That is, we have   f (A) =  r(A) if 

1 1/21  −1/2

β 0(t)β 1(t)

 =

  (1/2)t

(−1/2)t

  is satisfied. Solving this equation for  β 0(t) and  β 1(t), we obtain that

At = 2−

tI =

2−t

00 2−t

 if  t  is even, and  At = 2−

(t−

1)A =

2−t

2−(t−1)

0   −2−t

  if  t  is odd.

5.3.3 Decomposition Theorem for Matrices

The following theorem, called the primary decomposition theorem for linear transformations , showsthat every linear transformation on a complex finite-dimensional vector space is decomposed intopieces, each of which is a nilpotent matrix plus a scalar multiple of the identity matrix.

Theorem 5.10   Let  λ1, . . . ,  λ p  be the distinct eigenvalues of  A ∈  Cn×n. Let  mk   be the algebraic multiplicity of  λk, and let  V k  be the subspace of   Cn spanned by the generalized eigenvectors of   A

associated with  λk   for  k = 1, . . . ,  p. Then 

Cn = V 1 ⊕ · · · ⊕ V  p =  N (λ1I − A)n ⊕ · · · ⊕ N  (λ pI − A)n,

where each  V k   is invariant under   A  with  V k   = N (λkI − A)n and   dim V k   =  mk. Moreover, if   Vk

is a full-column-rank matrix whose columns span  V k, then   AVk   =  VkAk   for some   Ak  ∈  Cmk×mk

such that  λkI − Ak   is nilpotent for  k = 1, . . . ,  p.

Proof . We have seen in Section 5.3.1 that  V k   =  N (λkI − A)n and dim V k   =   mk   for each   k.If   x  ∈   V k, then (λkI − A)n(Ax) =   A(λkI − A)nx   =   A0 = 0, so each   V k   is invariant under   A.By Lemma 5.3, there exists an  Ak  ∈  C

mk×mk such that  AVk  =  VkAk   for each  k, so that (λkI −

Page 44: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 44/119

EE 580 — Linear Control Systems    8

A)nVk  =  Vk(λkI − Ak)n = 0. Since  Vk  has full column rank, this implies (λkI − Ak)n = 0. ThusλkI − Ak  is nilpotent for each  k . Now it remains to show  C

n = V 1 ⊕ · · · ⊕ V  p. Let V   = V 1 + · · · + V  p.Then   V   is a subspace of   Cn with dim V   ≤   n, where   n   =

  pk=1 mk   =

  pk=1 dim V k, and so the

proof is complete once we show dim V   =   n. Clearly   V   is invariant under   A. If   V   ∈   Cn×dimV 

is a full-column-rank matrix whose columns span  V  , then Lemma 5.3 implies that there exists a

matrix  AV   ∈ C

dimV ×dimV 

such that  AV  =  VAV   . Suppose (λI − A)

n

v  = 0 for some  λ ∈ C

 andv  ∈  V   . Then v  =  Vu  for some  u ∈  CdimV , giving (λI − A)nVu  =  V(λI − AV  )

nu = 0. Becausethe columns of  V  are linearly independent, we have (λI − AV  )

nu = 0. That is,  AV   has the sameeigenvalues, with the same multiplicities, as  A. This is possible only if dim V   = n.  

Corollary 5.11   Let  λ1, . . . ,  λ p  be the distinct eigenvalues of  A ∈ Cn×n. Then there is a basis  B

of  Cn consisting of generalized eigenvectors of  A   such that, with respect to  B, the matrix  A  has a block diagonal form  A  as follows:

A =

A1   0   · · ·   0

0 A2   · · ·   0...

  ...  . . .

  ...0 0   · · ·   A p

where    Ak  =

λk   ∗

. . .

0   λk

, k = 1, . . . , p.

Proof . For each  k, let V k  be the subspace of generalized eigenvectors of  A corresponding to  λk;let   Vk   be a full-column-rank matrix whose columns span   V k. Then by Theorem 5.10 we haveAVk   =  VkAk   for some matrix   Ak   ∈  C

dimV k×dimV k such that  λkI − Ak   is nilpotent. Choose aninvertible matrix Uk  ∈ C

dimV k×dimV k such that U−1k   (λkI − Ak)Uk   is strictly upper triangular (with

zero diagonal entries) for each  k. Then, because

AVkUk  =  VkAkUk  = VkUk(U−1k   AkUk) = VkUk(λkI − U−1

k   (λkI − Ak)Uk)

holds for each  k, the matrix  A =  P−1AP has the desired form with  P = [V1U1   · · ·   V pU p].  

5.3.4 Jordan Form

The primary decomposition theorem for linear transformations can be used to show that there isa basis of  Cn with respect to which the matrix of a linear transformation on the finite-dimensionalcomplex vector space contains zeros everywhere except possibly on the diagonal and the line directlyabove the diagonal. The following lemma is proved in [2, Lemma 8.40]:

Lemma 5.12   Let   N ∈  Cn×n be nilpotent. For each nonzero   v  ∈  C

n, let  m(v)  denote the largest nonnegative integer such that  Nm(v)v = 0. Then there exist  v1, . . . ,  vk  ∈ C

n such that 

(a)   {v1, Nv1, . . . , Nm(v1)v1, . . . , vk, Nvk, . . . , Nm(vk)vk}   is a basis of  Cn;

(b)   {Nm(v1)v1, . . . , Nm(vk)vk}  is a basis of  N (N).

For example, if  v1 = [1 0 0 0 0]T and  v2 = [0 0 0 1 0]T, then the nilpotent matrix

N =

0 0 0 0 01 0 0 0 00 1 0 0 00 0 0 0 00 0 0 1 0

Page 45: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 45/119

EE 580 — Linear Control Systems    9

has   m(v1) = 2 a n d   m(v2) = 1. Moreover,   {v1, Nv1, N2v1, v2, Nv2}   is a basis of   C5, and{N2v1, Nv2} is a basis of  N (N), which has dimension 2.

Given an A ∈ Cn×n, a basis of  Cn is called a Jordan basis   for A  if, with respect to this basis,  A

has a block diagonal matrix

J =

J1   0   · · ·   0

0 J2   · · ·   0...

  ...  . . .

  ...0 0   · · ·   Jm

,   where   Jk  =

λk   1 0

. . .   . . .. . .   1

0   λk

, k = 1, . . . , m.

This matrix   J   is called a   Jordan canonical form   of   A, and is both upper triangular and blockdiagonal. Each block  Jk  on the diagonal of  J   is called a  Jordan block . A Jordan block  Jk  may be just a 1-by-1 block consisting of just some eigenvalue; otherwise, the diagonal of   Jk   is filled withsome eigenvalue λk, the line directly above the diagonal is filled with 1’s, and all other entries are 0.If  p  denotes the number of distinct eigenvalues of  A, then the number  m  of Jordan blocks satisfies p ≤ m ≤ n.

Theorem 5.13  There exists a Jordan basis for every  A ∈ Cn×n.

Proof . First consider a nilpotent matrix N. Choose the vectors v1, . . . , vk given by Lemma 5.12.Then, for each  j, we have

N

Nm(vj)v j   · · ·   Nv j   v j

 =

Nm(vj)v j   · · ·   Nv j   v j

0 1 0. . .

  . . .. . .   1

0 0

.

Thus a Jordan form of  N  consists of  k  Jordan blocks. Now consider a general  A ∈ Cn×n. Let  λ1,

. . . , λ p  be the distinct eigenvalues of  A, with V 1, . . . ,  V  p  the corresponding subspaces of generalizedeigenvectors. Let  V1, . . . ,  V p  be full-column-rank matrices such that the columns of  V j   span  V  j

for each   j. Then by Theorem 5.10, we have  Cn =  V 1 ⊕ · · · ⊕ V  p   and  AV j   =  V jA j ,  j  = 1, . . . ,  p,where each   λ jI − A j   is nilpotent. Thus   A j  − λ jI   is nilpotent as well, and as we have just seenthere is a Jordan basis for it for each  j. Putting these bases together gives a basis of  Cn that is aJordan basis for  A.  

5.3.5 How to Obtain Jordan Canonical Form

We now know that a square matrix is always similar to its Jordan form, and the vectors that form aJordan basis are generalized eigenvectors of the matrix. A vector  v ∈ Cn is said to be a generalizedeigenvector of grade  k  of  A  associated with the eigenvalue  λ  if 

(A − λI)kv = 0 and (A − λI)k−1v = 0.

In this case, define

vk = v,

vk−1 = (A − λI)v = (A − λI)vk,

vk−2 = (A − λI)2v = (A − λI)vk−1,

...

v1 = (A − λI)k−1v = (A − λI)v2.

Page 46: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 46/119

EE 580 — Linear Control Systems    10

This set of vectors   {v1, . . . , vk}   is called a  chain of generalized eigenvectors   of length   k. It isclear that each  N (A − λI) j is a subspace of  N (λI − A) j+1. Moreover, we have that each  v j   is in

 N (A − λI) j but not in  N (A − λI) j−1. Therefore, the chain {v1, . . . , vk}  is linearly independent.The following example is from [4, Pages 39–41] and illustrates how to obtain a Jordan basis for  A:

1. Choose a repeated eigenvalue  λ  of algebraic multiplicity  m.   The task here is to find m linearly

independent generalized eigenvectors of  A  associated with  λ. This is achieved by searching chainsof generalized eigenvectors of various lengths.2. Compute the ranks of   (A − λI)k,   k   = 0,   1, ..., until the nullity of   (A − λI)k equals   m.

Such a   k   exists and is no more than   m. Let   N  j   denote the null space of (A − λI) j . For thepurpose of illustration, assume  n  = 10,   m  = 8,  k  = 4, and  v0   = dim N 0   = 0,   v1  = dim N 1   = 3,v2 = dim N 2  = 6,  v3 = dim N 3 = 7, and  v4 = dim N 4  = 8.

3. Find generalized eigenvectors of grade  k  associated with  λ  and generate chains of generalized eigenvectors based on them.   By assumption, we have  N 3   ⊂ N 4   and  v4 − v3  = 1, and so we canfind one and only one linearly independent vector   u, which is in   N 4   but not in   N 3, such that(A − λI)4u = 0 and (A − λI)3u = 0. From this  u, generate a chain of four generalized eigenvectorsas  u1 = (A − λI)3u,  u2 = (A − λI)2u,  u3 = (A − λI)u, and  u4 =  u.

4. Find generalized eigenvectors of grade  k − 1  associated with  λ  and generate chains of gener-

alized eigenvectors based on them.  By assumption, we have  N 2 ⊂ N 3  and  v3 − v2  = 1, and so thereis only one linearly independent vector which is in  N 3  but not in  N 2. The vector   u3  obtained inStep 3 above is such a vector, and the chain of generalized eigenvectors it generates is nothing but{u1, u2, u3}, which we already have.

...

5. Find generalized eigenvectors of grade   2  associated with   λ  and generate chains of generalized eigenvectors based on them.  Because v2−v1 = 3 by assumption, there are three linearly independentvectors that are in  N 2  but not in  N 1. Since u2  is one of them, we can find two vectors   v  and  w

such that   {u2, v, w}   is linearly independent and such that (A −  λI)2v   = 0 , (A −  λI)v   = 0,(A − λI)2w = 0, and (A − λI)w = 0. From  v  and  w, generate two chains of generalized eigenvalues

of length 2, say  {v1, v2}  and {w1, w2}.6. Find eigenvectors (i.e., generalized eigenvectors of grade  1) associated with  λ.  By assumption,

the number of vectors in  N 1   is equal to   v1 − v0   = 3. However,   u1,  v1, and   w1   are such vectors,and so there is no need to search for other vectors in  N 1.

7. Finally, form a linearly independent set of  m  generalized eigenvectors.  This completes thesearch of eight generalized eigenvectors of   A  associated with  λ. These vectors are  u1,  u2,  u3,  u4,v1,  v2,  w1, and  w2. The generalized eigenvectors generated this way are guaranteed to be linearlyindependent [4, Theorem 2-10]. Moreover, the generalized eigenvectors associated with differenteigenvalues are linearly independent due to Theorem 5.10.

8. Obtain a Jordan-form matrix. For each eigenvalue of  A, its geometric multiplicity is equal to the number of Jordan blocks associated with it, and its algebraic multiplicity is the sum of the 

orders of all Jordan blocks associated with it.   Let   V   = [u1   u2   u3   u4   v1   v2   w1   w2]. Because(A −  λI)u1   = 0 , (A −  λI)u2   =   u1, (A −  λI)u3   =   u2, (A −  λI)u4   =   u3, (A −  λI)v1   = 0,(A − λI)v2 =  v1, (A − λI)w1 = 0, and (A − λI)w2 =  w1, we have  AV =  VAV  , where

AV   =

λ   1 0 00   λ   1 00 0   λ   10 0 0   λ

λ   10   λ

λ   10   λ

.

Page 47: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 47/119

EE 580 — Linear Control Systems    11

Indeed, the number of Jordan blocks associated with  λ  is 3, which equals the geometric multiplicityof  λ. Also, these Jordan blocks form  AV   ∈ C

8×8, where 8 is the algebraic multiplicity of  λ.

5.3.6 An Example

Let  A ∈ C4×4 be as in Section 5.1.4 and Section 5.2.4:

A0 = I  =

1 0 0 00 1 0 00 0 1 00 0 0 1

,   A1 = A  =

1 0 1 00 0 0 0

−1 0 1 00 1 0 0

A2 =

0 0 2 00 0 0 0

−2 0 0 00 0 0 0

, . . . .

The roots of the characteristic polynomial det(sI−A) = s2[(s−1)2 +1] are 0, 0, 1+i, and 1−i. Therepeated eigenvalue λ  = 0 has algebraic multiplicity  m  = 2. We have rank(A− λI)0 = rank A0 = 4,rank(A − λI)1 = rank A = 3, and rank(A − λI)2 = rank A2 = 2, so  k  = 2 is the smallest numbersuch that dim N (A − λI)k = m  = 2. A generalized eigenvector  u such that (A − λI)ku =  A2u = 0and (A − λI)k−1u  =  Au  = 0 is then   u  = [0 1 0 0]T. From this   u, we generate a chain of  k  = 2generalized eigenvectors as  u2  = u  = [0 1 0 0]T and  u1  = (A − λI)u =  Au  = [0 0 0 1]T. Since theeigenvalues 1 + i  and 1 − i  give two more (generalized) eigenvectors—namely,   v1  = [1 0   − i  0]T

and   v2   = [1 0   − i  0]T, respectively—we have obtained a Jordan basis  {u1, u2, v1, v2}. LettingQ  = [u1   u2   v1   v2], which happens to be the same as in Example 5.2.4, gives a Jordan canonicalform  J  of  A  as

Q =

0 0 1 10 1 0 00 0   i   −i1 0 0 0

⇒   J =  Q−1AQ =

J1   0 0

0 J2   00   0   J3

 =

0 1 0 00 0 0 00 0 1 + i   00 0 0 1 − i

,

where the three Jordan blocks are  J1 = [ 0 10 0 ],  J2 = 1 + i, and  J3 = 1 − i.

References

[1] T. Kailath,  Linear Systems . Englewood Cliffs, NJ: Prentice-Hall, 1980.

[2] S. Axler, Linear Algebra Done Right , 2nd ed. New York, NY: Springer, 1997.

[3] R. A. Horn and C. R. Johnson, Matrix Analysis . Cambridge, UK: Cambridge University Press,1985.

[4] C.-T. Chen,  Linear System Theory and Design , 2nd ed. New York, NY: Oxford UniversityPress, 1984.

Page 48: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 48/119

EE 580 — Linear Control Systems

VI. State Transition Matrix 

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

6.1 Introduction

Typical signal spaces are (infinite-dimensional) vector spaces. Consider the space  C (J,Rn) of  Rn-valued continuous functions on an interval   J   = (a, b)  ⊂  R. For  f ,  g   ∈  C (J,Rn), define  f  + g   ∈

C (J,Rn

) to be the function satisfying

(f  +  g)(t) = f (t) + g(t) for all   t ∈ J.

Also, for  α ∈ R and  f  ∈ C (J,Rn), define αf  ∈ C (J,Rn) to be the function satisfying

(αf )(t) = αf (t) for all   t ∈ J.

Then it is straightforward to show that   C (J,Rn) is a vector space over the real field, where theorigin 0   ∈   C (J,Rn) is the function which is identically zero. Under this definition, a functionf   ∈   C (J,Rn) is a linear combination of functions   g1, . . . ,   gm   ∈   C (J,Rn) if there are   α1, . . . ,αm  ∈ R such that f (t) = (α1g1 + · · ·+αmgm)(t) for all t ∈ J . A nonempty finite subset {g1, . . . , gm}of   C (J,Rn) is then linearly independent if and only if (α1g1 + · · · + αmgm)(t) = 0 for all   t  ∈   J implies  α1 = · · · = αm = 0.

For simplicity, we will focus on LTI systems of the form  x(t) = Ax(t). However, all the theoremsup to Section 6.3.2 inclusive carry over to LTV systems of the form  x(t) = A(t)x(t) as long as  A(·)is continuous; see [1, Sections 2.3 & 2.4] and [2, Pages 50–73]. These results hold true even if  A(·)is piecewise continuous; a function is said to be  piecewise continuous  if, on each bounded intervalin its domain, it is continuous except p ossibly for a finite number of discontinuities. If   A(·) ispiecewise continuous (or piecewise constant in particular), then one can solve  x(t) = A(t)x(t) pieceby piece and then put together the results to obtain an overall solution.

6.2 Fundamental Matrices [3, Section 3.2.1]

6.2.1 Solution Space

Consider the homogeneous linear time-invariant system

x(t) = Ax(t), t ∈ R,   (6.1)

where   A   ∈   Rn×n is a constant matrix. We have seen that this system, subject to any initial

condition   x(t0) =  x0, has a unique solution  φ  on  R; that is, there exists a unique  φ  ∈  C (R,Rn)such that  φ(t) = Aφ(t) for all t ∈ R and such that φ(t0) = x0. It turns out that the set of solutionsto (6.1) is a finite-dimensional subspace of  C (R,Rn), which is infinite dimensional.

1

Page 49: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 49/119

EE 580 — Linear Control Systems    2

Theorem 6.1  The set of all solutions to  (6.1) forms an  n-dimensional linear subspace of  C (R,Rn).

Proof . Let   V   be the space of all solutions of (6.1) on   R. It is easy to verify that   V   is alinear subspace of  C (R,Rn). Choose a set of  n  linearly independent vectors  {x1, . . . , xn}  in  Rn. If t0  ∈  R, then there exist  n  solutions  φ1, . . . ,  φn   of (6.1) such that  φ1(t0) =  x1, . . . ,  φn(t0) =  xn.The linear independence of  {φ1, . . . ,φn} follows from that of {x1, . . . , xn}. It remains to show that

{φ1, . . . ,φn} spans V  . Let φ be any solution of (6.1) on R such that φ(t0) = x0. There exist uniquescalars  α1, . . . ,  αn  ∈  R  such that  x0  = α1x1 + · · · + αnxn, and thus that  ψ  = α1φ1 + · · · + αnφn

is a solution of (6.1) on  R  such that  ψ(t0) = x0. However, because of the uniqueness of a solutionsatisfying the given initial data, we have that  φ   =   α1φ1  + · · · + αnφn. This shows that everysolution of (6.1) is a linear combination of  φ1, . . . ,  φn, and hence that  {φ1, . . . ,φn}   is a basis of the solution space.  

6.2.2 Matrix Differential Equations

With  A ∈ Rn×n as before, consider the  matrix differential equation 

X(t) = AX(t).   (6.2)

If the columns of  X(t) ∈ Rn×n are denoted  x1(t), . . . ,  xn(t) ∈ R

n, then (6.2) is equivalent to

x1(t)...

xn(t)

=

A   · · ·   0...

  . . .  ...

0   · · ·   A

x1(t)...

xn(t)

,

so a matrix differential equation of the form (6.2) is a system of  n2 differential equations.

Definition 6.2   A set   {φ1, . . . ,φn} ⊂   C (R,Rn)   of   n   linearly independent solutions of   (6.1)   is called a   fundamental set of solutions  of   (6.1), and the  Rn×n-valued function 

Φ =φ1   · · ·   φn

is called a   fundamental matrix  of   (6.1).

Theorem 6.3  A fundamental matrix   Φ  of   (6.1)   satisfies the matrix equation   (6.2).

Proof . The result is immediate from

Φ(t) =φ1(t)   · · ·   φn(t)

 =

Aφ1(t)   · · ·   Aφn(t)

 =  Aφ1(t)   · · ·   φn(t)

 =  AΦ(t).  

Therefore, a fundamental matrix is a matrix-valued function whose columns span the space of solutions to the system of  n differential equations (6.1).

6.2.3 Characterizations of Fundamental Matrices

It can be shown that, if  Φ is a solution of the matrix differential equation (6.2), then

det Φ(t) = e(trA)(t−τ ) det Φ(τ ) (6.3)

for all t, τ  ∈ R. This result is called Abel’s formula  and proved in [1, Theorem 2.3.3]. Its immediateconsequence is that, since t, τ  are arbitrary, we have either det Φ(t) = 0 for all  t ∈ R or det Φ(t) = 0for any  t ∈  R. In fact, we have the following characterizations of fundamental matrices, which weobtain without using Abel’s formula.

Page 50: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 50/119

EE 580 — Linear Control Systems    3

Theorem 6.4   A solution   Φ   of the matrix differential equation   (6.2)   is a fundamental matrix of   (6.1)  if and only if  Φ(t)   is nonsingular for all   t ∈ R.

Proof . To show necessity by contradiction, suppose that   Φ   = [φ1   · · ·   φn] is a fundamentalmatrix for (6.1) and that det Φ(t0) = 0 for some   t0   ∈  R. Then, since   Φ(t0) is singular, the set{φ1(t0), . . . ,φn(t0)} ⊂   R

n is linearly dependent (over the real field), so that there are   α1, . . . ,

αn  ∈  R, not all zero, such that 

ni=1 αiφi(t0) = 0. Every linear combination of the columns of a

fundamental matrix is a solution of (6.1), and son

i=1 αiφi is a solution of (6.1). Due to the unique-ness of the solution, this, along with the initial condition

ni=1 αiφi(t0) = 0, implies that

ni=1 αiφi

is identically zero, which contradicts the fact that  φ1, . . . ,  φn  are linearly independent. Thus, weconclude that det Φ(t)   = 0 for all   t   ∈   R. To show sufficiency, let   Φ  be a solution of (6.2) andsuppose that det Φ(t) = 0 for all  t ∈ R. Then the columns of  Φ form a linearly independent set of vectors for all  t ∈ R. Hence,  Φ  is a fundamental matrix of (6.1).   .

For example, consider three  R3×3-valued functions

Φ1(t) =

1 2t t2

0 1   t0 0 1

,   Φ2(t) =

1   t t2

0 1   t0 0 1

,   Φ3(t) =

1   t t2

0 1   t0 0   t

, t ∈ R.

If we write  Φ1  = [φ1  φ2  φ3], where  φi ∈ C (R,R3) are given by  φ1(t) = [1 0 0]T,  φ2(t) = [2t 1 0]T,and  φ3(t) = [t2 t   1]T for all   t   ∈   R, then   {φ1,φ2,φ3}   is a linearly independent set in   C (R,R3)because

 3i=1 αiφi  being identically zero implies all  αi   equal 0. Similarly, the columns of  Φ2   are

linearly independent in  C (R,R3) and so are those of  Φ3. Thus,  Φ1,  Φ2, and  Φ3  are “potentially”fundamental matrices. We have det Φ1(t) = det Φ2(t) = 1 and det Φ3(t) =  t   for all   t  ∈  R. Sincedet Φ3(0) = 0, however, Theorem 6.4 tells us that  Φ3  is not a fundamental matrix of (6.1) or evenits time-varying version  x(t) =   A(t)x(t) (as the proof of Theorem 6.4 carries over to the time-varying case). On the other hand, since Φ1(t) and  Φ2(t) are nonsingular for all   t, it remains thatΦ1 and  Φ2 are fundamental matrices provided that they solve the matrix differential equation (6.2).Solving  Φ1(t) = AΦ1(t) for  A ∈ R3×3 yields that

A =  Φ1(t)Φ1(t)−1 =

0 2 2t

0 0 10 0 0

1   −2t t2

0 1   −t0 0 1

 =

0 2 0

0 0 10 0 0

,

so indeed  Φ1  is a fundamental matrix of the LTI system (6.1). However, since

Φ2(t)Φ2(t)−1 =

0 1 2t

0 0 10 0 0

1   −t   0

0 1   −t0 0 1

 =

0 1   t

0 0 10 0 0

is a function of  t, the matrix-valued function  Φ2   is not a fundamental matrix of (6.1). Neverthe-less,  Φ2  is a fundamental matrix of the LTV system  x(t) = A(t)x(t), where  A(t) =  Φ2(t)Φ2(t)−1.

Theorem 6.5   Let  Φ be a fundamental matrix of   (6.1). Then  Ψ  is a fundamental matrix of   (6.1)if and only if there exists a nonsingular matrix  P ∈ R

n×n such that  Ψ(t) = Φ(t)P  for all   t ∈ R.

Proof . Suppose P   is invertible and  Ψ(t) = Φ(t)P for all  t. Then it is easy to check that  Ψ  is asolution of (6.2). Moreover, since det Φ(t) = 0 for all   t  and since det P = 0, we have det Φ(t)P =det Φ(t)det P = 0 for all t. Thus by Theorem 6.4,  Ψ   is a fundamental matrix. Conversely, supposethat  Ψ   is a fundamental matrix. As  Φ(t)Φ(t)−1 = I  for all  t, the chain rule gives

ddt

ΦΦ−1

(t) = 0   ⇒   d

dt

Φ−1

(t) = −Φ(t)−1  d

dt

Φ

(t)Φ(t)−1 = −Φ(t)−1A

Page 51: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 51/119

EE 580 — Linear Control Systems    4

for all  t ∈ R. Using this equality, we obtain

ddt

Φ−1Ψ

(t) = Φ(t)−1  d

dt

Ψ

(t) +   ddt

Φ−1

(t)Ψ(t) = Φ(t)−1AΨ(t) − Φ(t)−1AΨ(t) = 0

for all   t. Hence   Φ(·)−1Ψ(·) is constant, which implies the existence of a matrix   P   such thatΦ(t)−1Ψ(t) = P  for all   t; since  Φ(t)−1 and  Ψ(t) are both nonsingular,  P   is nonsingular.  

6.3 State Transition Matrix [3, Sections 3.2.2 & 3.3.1]

6.3.1 Definitions

Solving the matrix differential equation (6.2) subject to the initial condition   X(t0) =  I  ∈  Rn×n is

equivalent to solving the initial-value problems

xi(t) = Axi(t), t ∈ R;   xi(t0) = ei  ∈ Rn,   (6.4)

separately over all   i  = 1, . . . ,  n, where   ei  denotes the   ith standard basis vector (i.e.,   ith columnof the identity matrix  I). That is, if  φ1, . . . ,  φn ∈ C (R,Rn) are the unique solutions to (6.4) such

that  φi(t0) =  ei  and  ˙φ(t) =  Aφ(t) for all   t ∈

 R and for each   i  = 1, . . . ,  n, then the fundamentalmatrix  Φ(·, t0) = [φ1(·)   · · ·  φn(·)] is the unique solution to the matrix differential equation

∂ 

∂tΦ(t, t0) = AΦ(t, t0)

subject toΦ(t0, t0) = I.

The  Rn×n-valued function Φ(·, t0) is called the  state transition matrix   of (6.1). Every  x0 ∈ Rn is a

linear combination of the standard basis vectors  e1, . . . ,  en   (i.e.,  x0 = Ix0), and so, as we alreadyknow, the solution to the system of differential equations (6.1) subject to the initial conditionx(t0) = x0  has a unique solution given by  φ(·) = Φ(·, t0)x0, which is the same linear combination

of the columns of the state transition matrix. Moreover, Φ(t, t0) is determined by the Peano-Bakerseries, which in the case of LTI systems reduces to the  matrix exponential  given by

Φ(t, t0) = eA(t−t0) =∞k=0

Ak(t − t0)k

k!

for all  t, t0 ∈ R with the convention that  A0 = I  and 0! = 1.

6.3.2 Properties of State Transition Matrix

Theorem 6.6   Let  Φ(t, t0)  be the state transition matrix of   (6.1). If  P  is invertible, then the state variable change  Px(t) = x(t)   leads to the   equivalent  state equation 

˙x(t) =  Ax(t), t ∈ R;   A =  P−1AP.   (6.5)

If  Φ(·, t0)  is the state transition matrix of   (6.1), then the state transition matrix of   (6.5)   is 

Φ(t, t0) = P−1Φ(t, t0)P, t, t0 ∈ R.   (6.6)

Proof . Equation (6.5) follows from x(t) = P−1x(t) and   ddt

P−1x

(t) = (P−1AP)P−1x(t), and

equation (6.6) from   ∂ ∂t

P−1Φ(t, t0)P

 = (P−1AP)P−1Φ(t, t0)P  and  P−1Φ(t0, t0)P =  I.  

Page 52: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 52/119

EE 580 — Linear Control Systems    5

Theorem 6.7   Let  Φ(·, t0)  be the state transition matrix of   (6.1). Then the following hold:

(a) If  Ψ  is any fundamental matrix of   (6.1), then  Φ(t, t0) = Ψ(t)Ψ(t0)−1  for all   t,   t0 ∈ R;

(b)   Φ(t, t0)  is nonsingular and  Φ(t, t0)−1 = Φ(t0, t)  for all  t,  t0 ∈ R;

(c)  Φ

(t, t0) = Φ

(t, s)Φ

(s, t0)  for all   t,  s,  t0 ∈R

; (semigroup property)

Proof . Part (a) follows from   ∂ ∂t

Ψ(t)Ψ(t0)−1

 =  AΨ(t)Ψ(t0)−1 and Ψ(t0)Ψ(t0)−1 = I. Choose

any fundamental matrix   Ψ. Then det Ψ(t)   = 0 for all   t, so (a) implies   Φ(t, t0) is nonsingu-lar for all   t,   t0   because det Φ(t, t0) = det

Ψ(t)Ψ(t0)−1

 = det Ψ(t)det Ψ(t0)−1 = 0. Moreover,

Φ(t, t0)−1 =

Ψ(t)Ψ(t0)−1−1

= Ψ(t0)Ψ(t)−1, so (b) holds. Finally, for any fundamental matrix Ψ,we have  Φ(t, t0) = Ψ(t)Ψ(t0)−1 =

Ψ(t)Ψ(s)−1

Ψ(s)Ψ(t0)−1

, so (c) holds.  

As an example, consider the homogeneous LTI system  x(t) = Ax(t) given by

x1(t)x2(t)x3(t)

 =

0 2 00 0 10 0 0

x1(t)x2(t)x3(t)

.

It follows readily from x3(t) = 0, x2(t) =   x3(t), and x1(t) = 2x2(t) that   φ1(t) = [ 1 0 0 ]T,φ2(t) = [2t  1 0]T, and  φ3(t) = [t2 t  1]T are solutions of the system among infinitely many others.Moreover,   {φ1,φ2,φ3}   is a linearly independent set in   C (R,R3). Thus   Ψ   = [φ1   φ2   φ3] is afundamental matrix of the system. Then by part (a) of Theorem 6.7 we have that the statetransition matrix of the system is

Φ(t, t0) = Ψ(t)Ψ(t0)−1 =

1 2t t2

0 1   t0 0 1

1   −2t0   t2

0

0 1   −t0

0 0 1

 =

1 2(t − t0) (t − t0)2

0 1   t − t0

0 0 1

.

Indeed, we have Φ(t, t0) = eA(t−t0) for all t, t0  ∈ R. As another example, consider the homogeneousLTV system  x(t) = A(t)x(t) given by

x1(t)

x2(t)x3(t)

 =

0 1   t

0 0 10 0 0

x1(t)

x2(t)x3(t)

.

As  φ1(t) = [1 0 0]T,  φ2(t) = [t  1 0]T, and  φ3(t) = [t2 t  1]T are linearly independent solutions tothis LTV system, the matrix-valued function  Ψ = [φ1  φ2  φ3] is a fundamental matrix of the LTVsystem (but not of any LTI system). Therefore, the state transition matrix of the LTV system is

Φ(t, t0) =

 Ψ(t)

Ψ(t0)

−1

=

1   t t2

0 1   t0 0 1

1   −t0   0

0 1   −t00 0 1

 =

1   t − t0   t(t − t0)

0 1   t − t00 0 1

.

That is,   ∂ ∂t Φ(t, t0) = A(t)Φ(t, t0) for all  t ∈ R with  Φ(t0, t0) = I.

6.3.3 Properties of Matrix Exponentials

Theorem 6.8   Let  A ∈ Rn×n. Then the following hold:

(a)   eAt1eAt2 = eA(t1+t2)  for all   t1,  t2 ∈ R;

Page 53: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 53/119

EE 580 — Linear Control Systems    6

(b)   AeAt = eAtA  for all  t ∈ R;

(c)

eAt−1

= e−At  for all  t ∈ R;

(d)   det eAt = e(trA)t  for all   t ∈ R;

(e) If  AB =  BA, then  eAteBt = e(A+B)t  for all  t ∈ R.

Proof . Part (a) holds because eAt1eAt2 =  Φ(t1, 0)Φ(0, −t2) =  Φ(t, −t2) = eA(t1+t2), where thesecond equality follows from Theorem 6.7(c). Part (b) follows from

A

  limm→∞

m

k=0

Aktk

k!

 = lim

m→∞

m

k=0

Ak+1tk

k!   =

  limm→∞

m

k=0

Aktk

k!

A.

Part (c) is immediate from Theorem 6.7(b). If   Ψ   is a fundamental matrix of (6.1), then Theo-rem 6.7(a) gives det eAt = det

Ψ(t)Ψ(0)−1

 = det Ψ(t)/ det Ψ(0), where the last equality follows

from 1 = det I  = det

Ψ(0)Ψ(0)−1

 = det Ψ(0)det Ψ(0)−1. Thus Abel’s formula (6.3) gives (d).Finally, if  AB =  BA, then  AiB j = B jAi and hence we have

  limm→∞

m

i=0

Aiti

i! lim

l→∞l

k=0

Bjtj

 j! = lim

l→∞l

k=0

Bjtj

 j!   lim

m→∞m

i=0

Aiti

i! ,

which implies (e).  

In general, matrices do not commute. For example, if 

A =

0 10 0

  and   B =

0 01 0

,

then  AB = BA. In this case,

eAt =

1   t0 1

, eBt =

1 0t   1

, e(A+B)t =

12 (et + e−t)   1

2 (et − e−t)12 (et − e−t)   1

2 (et + e−t)

,

so  eAteBt = e(A+B)t.

6.4 How to Determine Matrix Exponentials [3, Section 3.3.2]

We have already seen that the state transition matrix of a linear (time-invariant or time-varying)system can be obtained from a fundamental matrix, whose columns are linearly independent solu-tions to the given homogeneous system. Other methods to obtain the state transition matrix of anLTI system are summarized in this section.

6.4.1 Infinite Series Method

Given a matrix  A ∈ Rn×n, evaluate the partial sum

Sm(t) =mk=0

tkk!

Ak = I + tA +  t2

2 A + · · · +  tm

m!Am

for   m   = 0, 1, . . . . Then, since   Sm(t)   →   eAt uniformly on any bounded interval   J   in   R, it isguaranteed that  eAt ≈  Sm(t) for all   t ∈  J  and for all sufficiently large  m. If  A   is nilpotent, thenthe partial sum  Sm(t), with  m =  n − 1, gives a closed-form expression for  eAt. For example,

A =

0   α0 0

  ⇒   A2 =

0 00 0

  ⇒   eAt = I + tA =

1   αt0 1

.

Page 54: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 54/119

EE 580 — Linear Control Systems    7

6.4.2 Similarity Transformation Method

Diagonalizable Case . If  A ∈ Rn×n has n linearly independent eigenvectors v1, . . . ,  vn corresponding

to its (not necessarily distinct) eigenvalues   λ1, . . . ,   λn, then let   P   = [v1   · · ·   vn]. Then, sinceJ =  P−1AP = diag{λ1, . . . , λn}, Theorem 6.6 gives

eAt = PeJtP−1 = P

eλ1t

0. . .

0   eλnt

P−1.

General Case . Generate   n   linearly independent generalized eigenvectors   v1, . . . ,   vn   of   A   ∈Rn×n such that   P   = [v1   · · ·   vn] takes   A   into the Jordan canonical form   J   =   P−1AP   =diag{J0, . . . , Jm} ∈ Cn×n, where  J1, . . . ,  Jm  are Jordan blocks of varying dimensions. Then

eAt = PeJtP−1 = P

eJ1t 0. . .

0   eJmt

P−1.

A Jordan block is the sum of a diagonal matrix and a nilpotent matrix; that is,  Jk  =  Λk + Nk  with

Λk  =

λk   0   · · ·   00   λk   · · ·   0...

  ...  . . .

  ...0 0   · · ·   λk

and   Nk  =

0 1 0. . .

  . . .. . .   1

0 0

,

where  λk   is some eigenvalue of   A   for each   k. Let   Jk   ∈  Cnk×nk for each   k   = 1, . . . ,   m, so thatm

k=1 nk   =   n. Then we have   Nnkk   =   0, and so the series defining   eNkt terminates for each   k.

Moreover, direct computation yields  ΛkNk  =  NkΛk   for each  k. Therefore, Theorem 6.8(e) yields

eJkt = eΛkteNkt = eλkt

1   t t2/2  · · ·

  tnk−1/(nk −

1)!0 1   t   · · ·   tnk−2/(nk − 2)!0 0 1   · · ·   tnk−3/(nk − 3)!...

  ...  ...

  . . .  ...

0 0 0   · · ·   1

, k = 1, . . . , m.   (6.7)

Note that, if  λk   in (6.7) is complex, then  Euler’s formula  gives

eλkt = e(σk+iωk)t = eσkt(cos ωkt + i sin ωkt).

For example, if  α = 0, then

A =0 0

α   0

  ⇒   J =  P

−1

AP =0 1

0 0

,   P =0 1

α   0

  ⇒   eAt

= P1   t

0 1

P

−1

= 1 0

αt   1

.

6.4.3 Cayley-Hamilton Theorem Method

In view of the Cayley-Hamilton Theorem, there exist  β i(m; ·) ∈  C (R,R) for   i = 0, . . . ,  n − 1 andm = 1, 2, . . . such that, for A ∈ R

n×n,

eAt = limm→∞

mk=0

tk

k!Ak = lim

m→∞

n−1i=0

β i(m; t)Ai =n−1i=0

limm→∞

β i(m; t)Ai.

Page 55: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 55/119

EE 580 — Linear Control Systems    8

Then, letting  β i(t) = limm→∞ β i(m; t) for  i  = 0, . . . , n − 1 and  t ∈ R, we obtain

eAt =n−1i=0

β i(t)Ai, t ∈ R.   (6.8)

Thus one can determine  eAt by obtaining  β i(t) for all   i  and  t. Let  f (s) and  g(s) be two analyticfunctions (i.e., functions locally given by a convergent power series; e.g., polynomials, exponentialfunctions, trigonometric functions, logarithms, etc.). Let p(s) =

 p j=1(s − λ j)mk be the character-

istic polynomial of  A. Then  f (A) = g(A) if 

dkf 

dsk (λ j) =

 dkg

dsk(λ j), k = 0, . . . , m j − 1, k = 1, . . . , p,   (6.9)

where  p

 j=1 m j   =  n. (Proof . Equations (6.9) imply that f (s) − g(s) has  p(s) as a factor, so theresult follows from the Cayley-Hamilton Theorem.) Thus the terms  β i(t) in (6.8) can be determinedby letting

f (s) =  est and   g(s) = β 0(t) + β 1(t)s + · · · + β n−1(t)sn−1.

For example, if  A

 = [

 0 0

α  0 ], then the characteristic polynomial of  A

 is  p(s) = det(sI

−A

) = s

2

. Letf (s) = est and  g(s) = β 0(t) + β 1(t)s. Then, as  f (0) = g(0) and   ∂f ∂s (0) =   ∂g

∂s (0) imply  β 0(t) = 1 andβ 1(t) = t, we obtain that  eAt = g(A) = I + tA.

6.4.4 Laplace Transform Method

We know that eAt is the inverse Laplace transform of (sI−A)−1 for A ∈ Rn×n. The partial fraction

expansion of each entry in (sI − A)−1 gives

(sI − A)−1 =

 p j=1

mj−1k=0

k!

(s − λ j)k+1A jk ,

where   λ1, . . . ,   λ p   are the distinct eigenvalues of   A, with corresponding algebraic multiplicitiesm1, . . . ,  m p  such that

 p j=1 mi   =  n, and where each   A jk   ∈  C

n×n is a matrix of partial fractionexpansion coefficients. Taking the inverse Laplace transform gives

eAt =

 p j=1

mj−1k=0

tkeλjtA jk .   (6.10)

If some eigenvalues are complex, conjugate terms on the right side of (6.10) can be combined usingEuler’s formula to give a real representation. If   A jk   is nonzero, then   tkeλjt is called a   mode   of the system  x(t) =  Ax(t),   t  ∈  R. It is easily seen that  x(t)  →  0 as   t  → ∞  under arbitrary initialconditions if each mode of the system converges to zero (i.e., each  λ j  has negative real part). For

example, if  A = [ 0 0α  0 ], then  e

At

is the inverse Laplace transform of (sI − A)−1

=   1/s   0

α/s2 1/s

. Themodes of the system  x(t) = Ax(t) in this example are 1 and  t.

6.5 Discrete-Time Case

6.5.1 State Transition Matrix

Given an  A ∈ Rn×n, consider the discrete-time initial-value problem of solving

x(t + 1) = Ax(t), t =  t0, t0 + 1, . . . ;   x(t0) = x0.   (6.11)

Page 56: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 56/119

EE 580 — Linear Control Systems    9

The discrete-time state transition matrix is defined to be the unique solution   Φ(·, t0)  ∈  Rn×n of 

the matrix difference equation

Φ(t + 1, t0) = AΦ(t, t0), t =  t0, t0 + 1, . . . ,

subject toΦ(t

0, t

0) = I.

That is,   Φ(t, t0) =   At−t0 for all   t,   t0   ∈   Z   with   t   ≥   t0, and as we have seen earlier the uniquesolution to (6.11) is given by   x(t) =   Φ(t, t0)x0,   t   ≥   t0. Unlike the continuous-time case, thedifference equation (6.11) cannot be solved backward in time unless   A   is nonsingular. This isbecause, unless  A  is invertible, the discrete-time state transition matrix  Φ(t, t0) is not invertible.An immediate consequence is that the semigroup property holds only forward in time; that is,Φ(t, t0) = Φ(t, s)Φ(s, t0) for  t ≥ s ≥ t0. Due to time-invariance, we have  Φ(t, t0) = Φ(t − t0, 0).

6.5.2 How to Determine Powers of Matrices

Given an  A ∈ Rn×n, if  J =  P−1AP is a Jordan canonical form of  A, then we have

A

t

= (PJP

−1

)

t

= PJ

t

P

−1

.In particular, if   J  = diag{λ1, . . . , λn}, then   Φ(t, 0) =  At =  P diag{λt

1, . . . , λtn}P−1. Also, by the

Cayley-Hamilton Theorem there exist  β i(·),  i = 1, . . . , n − 1, such that

At =n−1i=0

β i(t)Ai, t = 0,   1, . . . .   (6.12)

Thus, as in the continuous-time case, one can determine  At by letting

f (s) = st and   g(s) = β 0(t) + β 1(t)s + · · · + β n−1(t)sn−1,

and then solving (6.9) for the terms  β i(t). Finally, one can use the fact that  Φ(·, 0) is the inverse  z-transform of  z(zI−A)−1 = (I−z−1A)−1 to determine At. If  λ1, . . . ,  λ p are the distinct eigenvalues

of  A, with corresponding algebraic multiplicities  m1, . . . , m p  such that

 pi=1 mi =  n, we have

z(zI − A)−1 = z

 p j=1

mj−1k=0

k!

(z − λ j)k+1A jk

and hence

At =

 p j=1

min{mj−1,t}k=0

t!

(t − k)!λt−k j   A jk ,

where   A jk   ∈  Cn×n are the matrices of partial fraction expansion coefficients. If   A jk   is nonzero,

then   t!(t−k)! λt−k

 j   is called a mode  of the system  x(t +1) = Ax(t). We have  x(t) → 0 under arbitrary

initial conditions if each mode converges to zero (i.e., each  λ j  has a magnitude less than 1).

References

[1] P. J. Antsaklis and A. N. Michel,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

[2] W. J. Rugh, Linear System Theory . Englewood Cliffs, NJ: Prentice-Hall, 1993.

[3] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

Page 57: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 57/119

EE 580 — Linear Control Systems

VII. Stability of Linear Systems

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

7.1 Definitions and Facts

We will see that internal stability of linear systems depends on the structural properties of thestate matrix  A ∈ R

n×n that governs the dynamics via  x =  Ax. The Jordan canonical form  J  of  A

reveals a complete picture of the structure of  A. However,  J  in general contains complex entries.Therefore, we need to extend our usual definitions of vector and matrix norms to complex vectorsand complex matrices, respectively. If   A ∈  Cm×n, then denoted by   A∗ ∈  Cn×m is the   conjugate transpose   of   A; that is,  A∗ is equal to  AT with its entries replaced by their complex conjugates.The Euclidean norm  ·  on  Cn is then defined by

x =√ 

x∗x = (|x1|2 + · · · + |xn|2)1/2

for  x = [x1 · · ·   xn]T ∈  Cn. It can be easily verified that x >  0 for all nonzero  x ∈ Rn, and thatαx  = |α|x  and x +  y ≤ x + y   for all   x,   y ∈  Cn. The spectral norm  ·  on  Cn×n isdefined as before by A = max{Ax : x = 1}  for all  A ∈ C

n×n. It can be shown that

A = max{√ λ : λ  is an eigenvalue of  A∗

A}= max{

√ λ : λ  is an eigenvalue of  AA∗} = A∗.

By definition, Ax ≤ Ax  for all  A ∈ Cn×n and  x ∈ Cn. Also, we have that A >  0 for allnonzero A ∈ Cn×n, and that αA = |α|A  and A + B ≤ A + B  for all  A,  B ∈ Cn×n andα ∈ C. Another important property of the spectral norm is that it is  submultiplicative ; that is,

AB ≤ AB

for  A,  B ∈ Cn×n. Another useful fact to remember is that we have

   t1

t0

x(t) dt ≤    t1

t0 x(t)

dt   and  

  t1

t0

A(t) dt ≤    t1

t0 A(t)

dt

for all  t0, t1 ∈ R  with  t0 ≤ t1  and for all  x(·) ∈ C (R,Rn) and  A(·) ∈ C (R,Rm×n).It is obvious that all the above definitions and properties are valid for   A ∈  Rn×n, for which

A∗ =   AT. If   A   =   AT ∈   Rn×n, then   A   is said to be   symmetric . For   A ∈   R

n×n, we shallwrite   A ≥   0   to mean that   A   is symmetric and   nonnegative definite   (i.e.,   xTAx ≥   0 for allx ∈ R

n). Nonnegative definite matrices are also called  positive semidefinite . In particular, we shallwrite  A >  0  to indicate that  A  is symmetric and  positive definite   (i.e.,   xTAx >  0 for all nonzerox ∈ R

n).  All the eigenvalues of a real symmetric matrix are real . (Proof . If  A ∈ Rn×n is symmetric,

1

Page 58: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 58/119

EE 580 — Linear Control Systems    2

and if   Av   =  λv   for some nonzero   v ∈  Cn, then  λ∗v2 = (λv)∗v   =   v∗A∗v  =   v∗Av   =   λv2,

so   λ∗ =   λ.) Also,   all the eigenvalues of a real symmetric, nonnegative definite (resp. positive definite) matrices are nonnegative (resp. positive). (Proof . If   A ≥   0   and   Av   =   λv   for somenonzero   v ∈   C

n, then   λv2 =   v∗Av   = ℜ[v]TAℜ[v] + ℑ[v]TAℑ[v] ≥   0; the case of   A   >   0   issimilar.) If  A ∈  Rn×n is symmetric, and if  λmin(A) and  λmax(A) denote the smallest and largest

eigenvalues of  A, respectively, then we have  λmin(A)x2

≤ x

T

Ax ≤ λmax(A)x2

for all  x ∈R

n

;that is,  λmin(A)I ≤ A ≤ λmax(A)I; in particular, if  A   is real and symmetric,  λmin(A) ≥ 0 (resp.λmin(A)   >   0) if and only if   A ≥   0   (resp.   A   >   0). A consequence is that a matrix   A ≥   0   isnonsingular if and only if  A >  0. It is readily seen that, for any rectangular matrix  M ∈  Rm×n,we have  MTM ≥ 0  and  MMT ≥ 0.

7.2 Internal Stability [1, Sections 4.3–4.6][2, Chapters 6 & 7]

Let  A ∈  Rn×n. In this section, we will study boundedness properties and asymptotic behavior of solutions of the homogeneous differential equation

x(t) = Ax(t), t ∈ R,   (7.1)

whose state transition matrix is denoted by  Φ(t, t0) ∈ Rn×n for t,  t0 ∈ R.

7.2.1 Lyapunov Stability

Definition 7.1   The system   (7.1)   is said to be   (uniformly) stable (in the sense of Lyapunov)   if there exists a  c > 0  such that 

x(t) ≤ cx(t0) for all  t,  t0 ∈ R  with   t ≥ t0  and for all  x(t0) ∈ R

n.

If a system is uniformly stable in the sense of Lyapunov, then one can guarantee the entire statetrajectory  x(t),  t ≥ t0, remains within an arbitrarily small neighborhood of the origin by ensuringthat the initial state   x(t0) is sufficiently close to the origin. That is, for every ε >  0, there existsa   δ >  0 such that x(t)  < ε   whenever   t ≥   t0   and x(t0)  < δ . Since we focus on linear time-invariant systems, this apparently weaker requirement is in fact equivalent to what Definition 7.1requires. In Definition 7.1, uniformity refers to the fact that the constant c  in the definition can bechosen independently of the initial time t0; this uniformity-in-time property is of course automaticfor time-invariant systems. The following result shows that Lyapunov stability is equivalent to theboundedness of the state transition matrix.

Lemma 7.2   The system   (7.1)   is uniformly stable in the sense of Lyapunov if and only if there exists a  c > 0  such that 

Φ(t, t0) ≤ c

 for all  t,  t0

 ∈R  with   t

≥t0.

Proof . Suppose that such a   c   exists. Then x(t)   = Φ(t, t0)x(t0) ≤ Φ(t, t0)x(t0) ≤cx(t0)   whenever  t ≥ t0, so sufficiency holds. Conversely, suppose that x(t) ≤ cx0  whenevert ≥   t0   and   x(t0) =   x0 ∈   Rn. By the definition of spectral norm, there exists an   x0 ∈   Rn suchthat x0  = 1 and Φ(t, t0)x0  = Φ(t, t0). Choosing such an   x0  and letting   x(t0) =   x0   givesΦ(t, t0) = Φ(t, t0)x0 = x(t) ≤ cx0 = c, so necessity holds.  

This result is valid for linear time-varying systems as well. Since Lyapunov stability is a propertyof the state matrix  A  alone, we shall say that the matrix  A  is stable to mean that the system (7.1)is stable.

Page 59: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 59/119

EE 580 — Linear Control Systems    3

Theorem 7.3   The following are equivalent:

(a) The matrix   A  is uniformly stable in the sense of Lyapunov;

(b) There exists a  c >  0  such that  eAτ  ≤ c  for all  τ  ≥ 0;

(c) All eigenvalues of   A   have negative or zero real parts, and the geometric multiplicity is the same as the algebraic multiplicity for each eigenvalue with zero real part.

Proof . By Lemma 7.2, the matrix  A is stable if and only if  Φ(t, t0) = eA(t−t0) is bounded over allt − t0 ∈ [0, ∞), so (a) and (b) are equivalent. Let  P  be a nonsingular matrix such that  J  =  P−1AP

is in the Jordan form. Then it follows from  eJτ  =   P−1eAτ P   and   eAτ  =   PeJτ P−1 that   eAτ  isbounded on [0, ∞) if and only if  eJτ  is bounded on [0, ∞). Every entry of  eJτ  is of the form βtkeλτ ,where β  is a positive number, k  a nonnegative integer, and  λ  an eigenvalue of  A. If  ℜ[λ] <  0, thenit is readily seen that  βtkeλτ  is bounded on [0, ∞). If  ℜ[λ] = 0, then  βtkeλτ  is bounded on [0, ∞)if and only if  k  = 0 (i.e., if and only if the size of each of the Jordan blocks associated with theeigenvalue  λ   is one-by-one). Since every Jordan block associated with  λ   is one-by-one if and onlyif  λ  has the same geometric and algebraic multiplicities, equivalence of (b) and (c) follows.  

For example, consider

A1 =

 0 1−1 0

  and   A2 =

0 1 0

0 0 00 0   −1

.

The two eigenvalues ±i  of  A1  are distinct and their real parts are nonpositive, so  A1   is Lyapunovstable. On the other hand, the eigenvalues of   A2  are 0 and −1, where only one Jordan block isassociated with the repeated eigenvalue 0, so  A2  is not Lyapunov stable. Indeed,

eA1τ  =

  cos τ    sin τ 

−sin τ    cos τ 

  and   eA2τ  =

1   τ    00 1 0

0 0   e−τ 

;

while eA1τ  = 1 ≤ 1 for all  τ  ≥ 0, we have eA2τ  → ∞  as  τ  → ∞.

7.2.2 Asymptotic Stability

Definition 7.4  The system   (7.1)  is said to be   (uniformly) asymptotically stable  if it is stable and if given any  c > 0  there exists a  T > 0  such that 

x(t) ≤ cx(t0)

 for all  t0 ∈ R, for all   t ≥ t0 + T , and for all  x(t0) ∈ Rn.

Definition 7.5   The system  (7.1) is said to be  (uniformly) exponentially stable  if there exist  c, λ >0  such that 

x(t) ≤ ce−λ(t−t0)x(t0) for all  t,  t0 ∈ R  with   t ≥ t0  and for all  x(t0) ∈ R

n.

What exponential stability requires is that, no matter what the initial state is, one can guaranteethe state  x(t) remains within an arbitrarily small neighborhood of the origin by ensuring that thetime span t − t0   is sufficiently large. That is, exponential stability guarantees asymptotic stability;

Page 60: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 60/119

EE 580 — Linear Control Systems    4

the converse is not true for general dynamical systems because exponential stability requires notonly that the state decays over time but also that the decay rate is exponential. However, for linear(time-varying and time-invariant) systems, uniform asymptotic stability and uniform exponentialstability are equivalent. Note that we require uniformity in time (i.e.,  T   in Definition 7.4 and  c, λin Definition 7.5 are independent of  t0), which is automatic for time-invariant systems.

Lemma 7.6   The following are equivalent:

(a) The system   (7.1)   is uniformly asymptotically stable;

(b) The system   (7.1)   is uniformly exponentially stable;

(c) There exist  c,  λ >  0  such that  Φ(t, t0) ≤ ce−λ(t−t0)  for all  t,   t0 ∈ R  with  t ≥ t0;

(d) There exists  β > 0  such that  t

t0Φ(t, τ ) dτ  ≤ β   for all   t,  t0 ∈ R  with   t ≥ t0.

Proof . We will show (a) ⇒ (b) ⇒ (c) ⇒ (d) ⇒ (a). Suppose the system is asymptotically stable.Then, there exists a  T > 0 such that x(t0 + T ) ≤ 2−1x(t0), x(t0 + 2T ) ≤ 2−1x(t0 + T ), . . . ,

for all  t0  and  x(t0). This gives that, for  t ∈ [t0 + kT, t0 + (k + 1)T ) and  k = 0, 1, . . . , we have

x(t) ≤ Φ(t, t0 + kT )x(t0 + kT ) ≤ Φ(t, t0 + kT )2−kx(t0).

Moreover, since the system is Lyapunov stable as well, Lemma 7.2 implies that there is a   c >  0such that Φ(t, t0 +  kT ) ≤  c   whenever   t ≥   t0 +  kT . Thus, for  t ∈   [t0 +  kT, t0 + (k + 1)T ) andk = 0, 1, ..., we have

x(t) ≤ c 2−kx(t0) ≤ 2c 2−(t−t0)/T x(t0) = 2cet−t0T 

  ln( 12

)x(t0),

so putting   λ   = (−1/T )ln(1/2) establishes that (a) implies (b). Next, suppose the system isexponentially stable, so that x(t) ≤  ce−λ(t−t0)x(t0)   whenever   t ≥ t0   and  x(t0) ∈  Rn. Choose

an  x0 ∈ R

n

such that x0 = 1 and Φ(t, t0)x0 = Φ(t, t0). Then letting  x(t0) = x0  gives

Φ(t, t0) = Φ(t, t0)x0 = x(t) ≤ ce−λ(t−t0)x0 = ce−λ(t−t0),

so (b) implies (c). Now, suppose Φ(t, t0) ≤ ce−λ(t−t0) whenever  t ≥ t0. Then we have   t

t0

Φ(t, τ ) dτ  ≤   t

t0

ce−λ(t−τ ) dτ   =  c

λ

1 − e−λ(t−t0)

≤   c

λ

whenever   t  ≥   t0. Thus putting   β   =   c/λ   establishes that (c) implies (d). Finally, suppose tt0

Φ(t, τ ) dτ  ≤  β   holds whenever   t ≥  t0. It is readily seen that t

t0∂ 

∂τ Φ(t, τ ) dτ   =  I − Φ(t, t0).

Also, term-by-term differentiation of the Peano-Baker series yields   ∂ ∂τ  Φ(t, τ ) = −Φ(t, τ )A. Hence

we have  Φ(t, t0) = I +  tt0 Φ(t, τ )A dτ , which leads to

Φ(t, t0) ≤ I +

   t

t0

Φ(t, τ )A dτ  ≤ 1 + β A,   (7.2)

which implies the system is Lyapunov stable. Moreover, for  t > t0, we have

Φ(t, t0) =  1

t − t0

   t

t0

Φ(t, t0) dτ  ≤   1

t − t0

   t

t0

Φ(t, τ )Φ(τ, t0) dτ  ≤  β (1 + β A)

t − t0,

Page 61: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 61/119

EE 580 — Linear Control Systems    5

where the last inequality follows from (d) and (7.2). For any   c >  0, putting  T   =  β (1 +  β A)/cgives Φ(t, t0) ≤  c   for   t ≥   t0 +  T , or x(t) ≤ Φ(t, t0)x(t0) ≤  cx(t0)   for   t ≥   t0 +  T   andx(t0) ∈ R

n. Thus (d) implies (a).  

This lemma is also valid for time-varying systems with bounded A(·). As in the case of Lyapunovstability, asymptotic stability of the system (7.1) is determined solely by the structure (i.e., the

Jordan form) of  A. Therefore, we shall speak of asymptotic stability and exponential stability of the matrix  A.

Theorem 7.7   The following are equivalent:

(a) The matrix   A   is uniformly asymptotically stable;

(b) The matrix   A   is uniformly exponentially stable;

(c) There exist  c,  λ >  0  such that  eAτ  ≤ ce−λτ   for all  τ  ≥ 0;

(d) There exists  β > 0  such that  ∞

0   eAτ dτ  ≤ β ;

(e) All eigenvalues of  A  have negative real parts;

(f) There exists a  Q >  0  such that  ATQ + QA <  0.   (Lyapunov Inequality)

Proof . The equivalence of (a), (b), (c), and (d) is immediate from Lemma 7.6. To show (a)and (e) are equivalent, let  P  be a nonsingular matrix such that  J  =  P−1AP is in the Jordan form.It follows from  eJτ  = P−1eAτ P and  eAτ  = PeJτ P−1 that eJτ   is bounded and converges to zeroas  τ  → ∞   if and only if  eAτ    is bounded and converges to zero as  τ  → ∞. In particular, Sinceevery entry of  eJτ  is of the form βtkeλτ , where β   is a positive number,  k  a nonnegative integer, andλ  an eigenvalue of  A, we obtain that eAτ  → 0 if and only if all eigenvalues of  A  have negativereal parts. Thus (a) and (e) are equivalent.

We will complete the proof by showing that (c) and (f) are equivalent. Suppose eAτ  ≤ ce−λτ 

for all  τ  ≥ 0. LetQ =

   ∞0

eATτ eAτ  dτ.

SinceeA

Tτ  =

(eAτ )T = eAτ , we have

   ∞0

eATτ eAτ 

dτ  ≤   ∞

0

eATτ eAτ  dτ  ≤

   ∞0

c2e−2λτ  dτ  =  c2

2λ,

so  Q is well-defined. Using the fact that   ddτ e

Aτ  = AeAτ  = eAτ A, we obtain

xT

ATQ + QA

x =

   ∞0

xT

 d

dτ  eA

Tτ 

eAτ  + eA

Tτ 

 d

dτ  eAτ 

x dτ 

=   ∞

0

d

dτ eAτ x2 dτ  = eAτ x2

∞τ =0

= −x2

for all  x ∈ Rn, so  ATQ + QA = −I <  0. Now, to show (c) implies (f), it remains to prove  Q >  0.

By definition,  Q =  QT. Also, since

d

dτ 

xTeA

Tτ eAτ x

 =  xTeATτ 

AT + A

eAτ x ≥ −AT + AeAτ x

2 ≥ −2AxTeATτ eAτ x

Page 62: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 62/119

EE 580 — Linear Control Systems    6

for all  x ∈ Rn, we obtain

xTQx ≥ −   1

2A   ∞

0

d

dτ 

xTeA

Tτ eAτ x

dτ  ≥ −   1

2A   ∞

0

d

dτ eAτ x2 dτ  =

  1

2Ax2

for all  x

 ∈ R

n. That is,  Q

 ≥ (1/2

A

)I  >  0. This establishes the implication from (c) to (f ). To

show the converse, suppose   Q   >   0  and   ATQ +  QA   <   0. Let   η   =  λmin(Q),   ρ   =  λmax(Q), andν  = λmin(−ATQ − QA). Then

d

dt

x(t)TQx(t)

 =  x(t)T(ATQ + QA)x(t) ≤ −ν x(t)2 and   x(t)TQx(t) ≤ ρx(t)2

for  t ≥ t0, so thatd

dt

x(t)TQx(t)

≤ −ν 

ρx(t)TQx(t)

for   t ≥   t0. If we put   V  (t) =   x(t)TQx(t), then this implies after integrating from   t0   to   t   that

V  (t) ≤ e−νρ

(t−t0)V  (t0), and hence that

x(t)TQx(t) ≤ e−νρ (t−t0)x(t0)TQx(t0)

for  t ≥ t0, which gives

x(t)2 ≤  1

ηx(t)TQx(t) ≤   1

ηe−

νρ

(t−t0)x(t0)TQx(t0) ≤  ρ

ηe−

νρ

(t−t0)x(t0)2

for  t ≥ t0. Thus condition (c) holds with  c = 

ρ/η  and  λ  =  ν /2ρ. This completes the proof.  

For example, while none of the matrices   A1   and   A2   considered at the end of Section 7.2.1 isasymptotically stable, the matrix  A  =

−1  −1

1   −1

 is asymptotically stable (and hence exponentially

stable) because all its eigenvalues −1 ± i   have negative real parts. Also, with  Q   =   I, we haveATQ + QA =  AT + A =

−2 00   −2  <  0  and  Q = [ 1 0

0 1 ] >  0, so the Lyapunov inequality is satisfied.

In the proof of Theorem 7.7, a function of the form  V  (t) =  x(t)TQx(t) is used to show that thefeasibility of the Lyapunov inequality implies asymptotic stability. Such a function defines, in asense, the “total energy” of the system because there are continuous, strictly increasing functionsα(·) and β (·) with α(0) = β (0) = 0 such that  α(x(t)) ≤ V  (t) ≤ β (x(t)). Thus (it can be shownthat) the asymptotic stability of the system is guaranteed if there is a continuous, strictly increasingfunction γ (·) with  γ (0) = 0 such that  V  (t) ≤ −γ (x(t)). Such a “total energy” function that canbe used to show the asymptotic stability of a dynamical system is called a  Lyapunov function .As shown in the proof of Theorem 7.7, the feasibility of the Lyapunov inequality guarantees theexistence of a Lyapunov function and hence the asymptotic stability of the state matrix  A. Indeed,

we have  eAτ  =

e−τ  cos τ   −e−τ  sin τ e−τ  sin τ e−τ  cos τ 

, and so eAτ  → 0 as  τ  → ∞.

7.3 Input-Output Stability [1, Section 4.7][2, Chapter 12]

In this section, we will study the boundedness property of the zero-state response of linear systems

x(t) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t).(7.3)

The input-output behavior of (7.3), under the zero initial state, is specified by the impulse responsematrix  H(t, t0) = CΦ(t, t0)B + Dδ (t − t0) for  t ≥ t0, where  δ  is the Dirac delta function.

Page 63: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 63/119

EE 580 — Linear Control Systems    7

Definition 7.8  The system   (7.3)  is said to be   (uniformly) bounded-input bounded-output stable(or   BIBO stable) if there exists an  η > 0  such that, whenever  x(t0) = 0, we have 

y(t) ≤ η supτ ≥t0

u(τ )

 for all  t,  t0 ∈ R  with   t ≥ t0   and for all (piecewise continuous)  u  with  supτ ≥t0 u(τ ) < ∞.

Here, supτ ≥t0 u(τ )  denotes the supremum   (or least upper bound ) of the set {u(τ ) :  τ  ≥ t0}.That is, if  r  = supτ ≥t0 u(τ ), then we have that u(τ ) ≤ r for all τ  ≥ t0 and that r ≤ r′ wheneveru(τ ) ≤  r′ for all   τ  ≥   t0. Thus, the output of a BIBO stable system is bounded whenever theinput signal is bounded. Moreover, the uniformity requirement says that the single constant  η   inDefinition 7.8 works over all initial times   t0. Of course, this uniformity requirement is automaticand hence redundant for time-invariant systems.

Lemma 7.9   The system  (7.3) is uniformly bounded-input bounded-output stable if and only if there exists a  ρ > 0  such that 

   t

t0 H(t, τ ) dτ  ≤ ρ

 for all  t,  t0 ∈ R  with   t ≥ t0.

Proof . Suppose t

t0H(t, τ )dτ  ≤ ρ  for t ≥ t0. If  u  is bounded and  x(t0) = 0, we have

y(t) =

   t

t0

H(t, τ )u(τ ) dτ 

 ≤   t

t0

H(t, τ )u(τ ) dτ 

≤   t

t0

H(t, τ ) dτ 

 supτ ≥t0

u(τ )

≤ ρ supτ ≥t0

u(τ )

for t ≥ t0, and so putting η  =  ρ  yields the sufficiency part of the desired result. To show necessity,we will first consider the SISO case where  H(t, τ ) = h(t, τ ) and  y(t) =

 tt0

h(t, τ )u(τ ) dτ   whenever

x(t0) = 0 and t ≥ t0. Suppose the system is BIBO stable and yet no ρ > 0 satisfies t

t0|h(t, τ )|dτ  ≤ ρ

for all   t,   t0 ∈  R  with   t ≥  t0; that is, for each  ρ >  0, there exist   tρ   and  sρ   with   tρ  > sρ   such that tρsρ

|h(tρ, τ )| dτ > ρ. Define the input signal  u(t),  t ≥ sρ, by

u(t) =

1 if  t ∈ [sρ, tρ) and  h(tρ, t) >  0;

−1 if  t ∈ [sρ, tρ) and  h(tρ, t) <  0;

0 otherwise.

Clearly, this  u   is bounded with supτ ≥sρ |u(τ )| = 1. However, when  x(sρ) = 0, we have

|y(tρ)| =

   tρ

h(tρ, τ )u(τ ) dτ 

 =   tρ

|h(tρ, τ )| dτ > ρ =  ρ  supτ ≥sρ

|u(τ )|.

Since this holds for each   ρ >  0, there does not exist an  η >  0 such that |y(t)| ≤  η supτ ≥sρ |u(τ )|for all   t,   t0 ∈  R  with   t ≥   t0. This contradicts the BIBO stability assumption, and hence provesthe necessity part of the result for SISO cases. It remains to generalize necessity to MIMO cases.

Page 64: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 64/119

EE 580 — Linear Control Systems    8

Suppose no  ρ > 0 satisfies t

t0H(t, τ )dτ  ≤  ρ  for all   t,   t0 ∈  R  with  t ≥ t0. Then, for each  η > 0,

there exist iη, jη, sη, and tη  with tη  > sη  such that tη

sη|hiη jη(tη, τ )| dτ > η, where hij  denotes entry

(i, j) of  H. By mimicking the SISO case, we can define a bounded u  with supτ ≥sη u(τ )  = 1, sothat y(tη) > η . This way, necessity is established for MIMO cases as well.  

Theorem 7.10 If the system  (7.3) is uniformly asymptotically stable, then it is uniformly bounded-input bounded-output stable as well.

Proof . By Lemma 7.6 we have that there are constants  c,  λ > 0 such that   t

t0

H(t, τ ) dτ  ≤   t

t0

CΦ(t, τ )B dτ  + D ≤   t

t0

ce−λ(t−τ ) dτ 

CB + D

≤ (c/λ)CB + Dwhenever  t ≥ t0. Thus Lemma 7.9 with  ρ = (c/λ)CB + D  establishes the result.  

The above results—namely, Lemma 7.9 and Theorem 7.10—are also valid for LTV systems.Because   y(t) =

 tt0

CΦ(t, τ )Bu(τ ) dτ  + Du(t),   t ≥ t0, whenever  x(t0) = 0, and  Du(·) is boundedwhenever   u(

·) is bounded, BIBO stability is independent of the “direct transmission” matrix   D.

Thus we will speak of the BIBO stabilty of triples (A, B, C). According to Theorem 7.10, however,if  A  is asymptotically stable, then the triple (A, B, C) is BIBO stable regardless of  B or  C.

Theorem 7.11   Let  A ∈ Rn×n,  B ∈ R

n×m, and  C ∈ Rl×n. The following are equivalent:

(a) The triple  (A, B, C)   is uniformly bounded-input bounded-output stable;

(b) There exists  ρ >  0  such that  ∞

0

CeAτ Bdτ  ≤ ρ;

(c) All the poles of each entry of  C(sI − A)−1B  have negative real parts.

Proof . Equivalence of (a) and (b) is due to Lemma 7.9, so it remains to show that (b) and (c)are equivalent. Each entry of  C(sI

−A)−1B is a strictly proper rational function (i.e., the degree

of the numerator polynomial is strictly less than the degree of the denominator polynomial), andhence can be expanded by partial fraction expansion into a sum of a finite number of terms of theform  β/(s − λ)k, where  λ   is a pole. Consequently, each entry of the impulse response matrix is asum of a finite number of terms of the form  βτ k−1eλτ , which is integrable if and only if  λ  has anegative real part. Hence we conclude that conditions (b) and (c) are equivalent.  

For example, consider  x(t) = Ax(t) + Bu(t) and  y (t) = Cx(t), where

A =

−1 00 1

,   B =

11

,   and   C =

1 0

.

Since the eigenvalues of  A are ±1 and one of them has positive real part, the system is internallyunstable. However, the transfer function of the system is

 h(s) = C(sI − A)−1B = 1/(s + 1), so the

system is BIBO stable. While the system internally has two modes (namely,  e−τ 

and eτ 

), only thestable mode e−τ  is “visible” from outside; that is, the unstable mode  eτ  does not manifest itself asa pole of the transfer function from  u  to  y. Thus, such a system cannot be stabilizable with anyfeedback controller that generates the control input  u  based on the measurement  y. In this case,one might want to replace either the input matrix   B   (i.e., the actuator) or the output matrix   C

(i.e., the sensor) to make the unstable mode appear as a pole of the transfer function. For instance,if we replace  C  with C = [0 1], then the transfer function becomes 1/(s − 1) and it is possible tostabilize the system via feedback control; in this case, the pole at  s = −1, which does not appearin the transfer function, is asymptotically stable and so it does not affect the closed-loop stability.

Page 65: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 65/119

EE 580 — Linear Control Systems    9

7.4 Stability of Discrete-Time Linear Systems [1, Section 4.8][3, Section 6.10]

Now, let us consider the internal stability and input-output stability of discrete-time linear time-invariant systems of the form

x(t + 1) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t).  (7.4)

The results and their proofs are analogous to the continuous-time cases, and thus we will presentthem without proof.

Definition 7.12   The system  (7.4) is said to be  (uniformly) exponentially stable  if there exist  c > 0and  λ ∈ (0, 1)  such that, whenever  u(t) = 0  for all  t ≥ t0,

x(t) ≤ cλt−t0x(t0)

 for all  t0,  t ∈ Z  with   t ≥ t0  and for all  x(t0) ∈ Rn.

Theorem 7.13   Let  A ∈ Rn×

n. The following are equivalent:

(a) The discrete-time linear time-invariant system   (7.4)   is uniformly exponentially stable;

(b) There exist  c > 0  and  λ ∈ (0, 1)  such that  Ak ≤ cλk  for all  k = 0,  1, . . . ;

(c) The magnitude of every eigenvalue of  A   is less than one;

(d) There exists a  Q >  0  such that  ATQA − Q <  0. (discrete-time Lyapunov inequality)

Definition 7.14   The system   (7.4) is said to be  (uniformly) bounded-input bounded-output stableif there exists an  η > 0  such that, whenever  x(t0) = 0, we have 

y(t) ≤ η supτ ≥t0

u(τ )

 for all  t,  t0 ∈ Z  with   t ≥ t0  and for all  u  with  supτ ≥t0 u(τ ) < ∞.

Theorem 7.15   Let  A ∈ Rn×n,  B ∈ R

n×m, and  C ∈ Rl×n. The following are equivalent:

(a) The system   (7.4)   is uniformly bounded-input bounded-output stable;

(b) There exists  ρ >  0  such that  ∞

k=0 CAkB ≤ ρ;

(c) All the poles of each entry of  C(zI − A)−1B  have magnitude less than one.

References

[1] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[2] W. J. Rugh, Linear System Theory . Englewood Cliffs, NJ: Prentice-Hall, 1993.

[3] P. J. Antsaklis and A. N. Michel,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

Page 66: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 66/119

EE 580 — Linear Control Systems

VIII. Controllability of Linear Systems

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

8.1 Introduction

Analogous to the fact that matrices are linear maps between finite-dimensional vector spaces, lineardynamical systems are linear operators between infinite-dimensional signal spaces. If  T  :  C (R,Rm) →

C (R,Rl

) is a linear operator, for instance, then the  null space   (or  kernel ) of  T   is defined as

 N (T ) = {v ∈ C (R,Rm :  T v = 0},

and the range   (or  image ) of  T   is defined as

R(T ) = {w ∈ C (R,Rl) :   w =  T v,   v ∈ C (R,Rm)}.

Clearly,  N (T ) and  R(T ) are linear subspaces of  C (R,Rm) and  C (R, Rl), respectively. These sub-spaces for other types of linear operators, like   T  :  C (R, Rm)   →   R

n or   T  :  Rn →   C (R,Rl), aredefined in the obvious way. Again, our notational convention is as follows: If   u  ∈  C (R,Rm) andT  :  C (R,Rm) → C (R,Rl), then, while   u =  u(·) ∈ C (R, Rm) and T u = (T u)(·) ∈ C (R,Rl), we have

u(t) ∈ R

m

and (T u)(t) ∈ R

l

for each  t ∈ R.Consider the state equation

x(t) =  Ax(t) + Bu(t),   (8.1)

where   A  ∈  Rn×n and   B  ∈   R

n×m are constant matrices. We are concerned with the problem of steering the state   x   of the system (8.1) to a desired value by generating an appropriate controlinput signal   u. This is a “controllability” problem. Such a problem arises, for instance, in track-seeking control of a hard disk drive, where the magnetic head of the hard disk drive needs to besteered from one position to another within a given time span. (Once the magnetic head has beenpositioned on or near the desired track, a track-following control mode would kick in to maintainthe head on the desired track to read the data on the disk; this is a “stability” problem.)

8.2 Reachable States

Definition 8.1   A state   x1  ∈  Rn is said to be  reachable  (or  controllable from the origin) if there 

exists an input   u  that transfers the state   x   from   x(0) = 0   to   x(T ) =  x1   for some  T  ∈ (0, ∞).

For  t0 ≤ t1, let  Lr(t0, t1) :  C (R,Rm) → Rn be the linear operator defined by

Lr(t0, t1)u =

   t1

t0

eA(t1−τ )Bu(τ ) dτ 

1

Page 67: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 67/119

EE 580 — Linear Control Systems    2

for   u ∈  C (R,Rm). We have   x(t) =  eAtx(0) + Lr(0, t)u   for all   t ∈  R, so a state   x1   is reachable if there exists an input   u   such that   x1  = Lr(0, T )u  for some  T   ∈  (0, ∞). If we denote the set of allreachable states by  Rr, then  Rr  constitutes a linear subspace of the state space  R

n given by

Rr  = {x1 ∈ Rn :   x1 ∈ R(Lr(0, T )),  0 < T < ∞}.

That is,   x1   is reachable if and only if   x1 ∈ Rr.

Definition 8.2   The  reachability Gramian  of the system   (8.1)  is the matrix 

Wr(t0, t1) =

   t1

t0

eA(t1−τ )BBTeA

T(t1−τ ) dτ  ∈ Rn×n, t0 ≤ t1.

Clearly, we have   Wr(t0, t1) ≥ 0  for all  t0,  t1 ∈ R with  t0 ≤ t1.

Lemma 8.3   R(Lr(t0, t1)) = R(Wr(t0, t1))  for all  t0,   t1 ∈ R  with  t0 ≤ t1.

Proof . For notational convenience, let us drop the arguments  t0   and   t1   whenever there is no

ambiguity. To show R(Wr) ⊂ R(Lr), choose an  x1 ∈ R(Wr). Then  x1 =  Wrη1 for some  η1 ∈ Rn.If we put   u(τ ) =  B

TeAT(t1−τ )η1, then

Lru =

   t1

t0

eA(t1−τ )Bu(τ ) dτ  =

   t1

t0

eA(t1−τ )BBTeA

T(t1−τ )η1 dτ  =  Wrη1  =  x1,

so   x1   ∈ R(Lr). This establishes the desired inclusion. Conversely, to show the reverse inclusionR(Lr)  ⊂ R(Wr), choose an   x1   ∈ R(Lr), so that there is a   u  ∈   C (R,Rm) satisfying   Lru  =   x1.Since  Wr  =  W

Tr , we have R(Wr)⊕N (Wr) = R

n, so there is a unique pair (xr, xn) of  xr  ∈ R(Wr)and   xn ∈ N (Wr) such that   x1  =  xr + xn. Since   Wrxn  = 0, we have   x

TnWrxn  = 0. Because

xTnWrxn  =   t1t0

xTneA

(t1−τ 

)BBTeAT

(t1−τ 

)xn dτ  =   t1t0

BTe

AT

(t1−τ 

)xn2

dτ,

whereB

TeAT(t1−τ )xn

2is continuous in τ  and nonnegative, this implies that   B

TeAT(t1−τ )xn  = 0

for all   τ   ∈   [t0, t1]. On the other hand, since   Wr   =   WTr  , we have   R(Wr)⊥ =   N (Wr), and so

xTr  xn  = 0, which implies

xTn xn  =  x

T1 xn  = (Lru)T

xn =

   t1

t0

u(τ )TB

TeAT(t1−τ )xn dτ.

This, along with   BTeA

T(t1−τ )xn   = 0 for all   τ   ∈   [t0, t1], implies that   xn   = 0, so that   x1   =   xr   ∈R(Wr). This establishes the reverse inclusion and hence completes the proof.  

Theorem 8.4   A state   x1 ∈ Rn is reachable if and only if 

x1 ∈ R(Wr(0, T )) (8.2)

 for some  T  ∈ (0, ∞). Moreover, if   (8.2) holds and if  η1 ∈ Rn is any vector such that   Wr(0, T )η1 =

x1, then an input   u  that transfers the state   x   from   x(0) = 0  to   x(T ) =  x1  is given by 

u(t) =  BTeA

T(T −t)η1, t ∈ [0, T ).

Page 68: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 68/119

EE 580 — Linear Control Systems    3

Proof . The result is immediate from Lemma 8.3 and the first half of its proof.  

As an example, consider

A =

1 00 1

  and   B =

10

,

which give the reachability Gramian

Wr(0, T ) =

   T 

0

eT −τ  0

0   eT −τ 

10

1 0

eT −τ  00   eT −τ 

 dτ  =

(e2T  − 1)/2 0

0 0

.

It is readily seen that R(Wr(0, T )) = span{[1 0]T} for T > 0. Let  x1 = [1 0]T. Then  x1 is reachablebecause   x1  ∈ R(Wr(0, T )), and hence, for every  T >  0, there is an input signal   u  that steers thestate of the system  x(t) =  Ax(t) + Bu(t) from   x(0) = 0 to   x(T ) =  x1. Since  η1 = [2/(e2T  − 1) 0]T,among infinitely many others, solve   Wr(0, T )η1 =  x1, such an input is given by

x1 =

10

  ⇒   u(t) =

1 0

eT −t 00   eT −t

2/(e2T  − 1)

0

 =

  2eT −t

e2T  − 1, t ∈ [0, T ).

Indeed, this   u  satisfies  Lr(0, T )u = T 

0   eA(T −τ )Bu(τ ) dτ   =  x1.

8.3 Controllable States

Definition 8.5   A state   x0  ∈  Rn is said to be  controllable   (or  controllable to the origin) if there 

exists an input   u  that transfers the state   x   from   x(0) =  x0   to   x(T ) = 0  for some  T  ∈ (0, ∞).

For  t0 ≤ t1, let  Lc(t0, t1) : C (R, Rm) → Rn be the linear operator defined by

Lc(t0, t1)u =

   t1

t0

eA(t0−τ )Bu(τ ) dτ 

for   u   ∈   C (R, Rm). A state   x0   is controllable if there exists an input   u   such that   −eAT x0   =Lr(0, T )u, or  −x0   =  Lc(0, T )u, for some  T   ∈  (0, ∞). Thus, the set of all controllable states is alinear subspace of the state space  Rn given by

Rc = {x0 ∈ Rn :   x0 ∈ R

Lc(0, T )

,  0 < T < ∞}.

Definition 8.6   The   controllability Gramian  of the system   (8.1)  is the matrix 

Wc(t0, t1) =

   t1

t0

eA(t0−τ )BBTeA

T(t0−τ ) dτ  ∈ Rn×n, t0 ≤ t1.

Clearly, we have   Wr(t0, t1) ≥ 0 for all  t0, t1 ∈ R with  t0 ≤ t1. The controllability Gramian and

the reachability Gramian are related via the identity

Wr(t0, t1) =  Φ(t1, t0)Wc(t0, t1)Φ(t1, t0)T.   (8.3)

Lemma 8.7   R

Lc(t0, t1)

 = R(Wc(t0, t1))  for all   t0,  t1 ∈ R  with   t0 ≤ t1.

Proof . The proof is virtually identical to that of Lemma 8.3.  

Page 69: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 69/119

EE 580 — Linear Control Systems    4

Theorem 8.8   A state   x0 ∈ Rn is controllable if and only if 

x0 ∈ R(Wc(0, T )) (8.4)

 for some  T  ∈ (0, ∞). If   (8.4)  holds and if   η0 ∈ Rn is such that   Wc(0, T )η0 =  x0, then an input   u

that transfers the state   x   from   x(0) = x0   to   x(T ) = 0  is given by 

u(t) = −BTe−A

Ttη0, t ∈ [0, T ).

Proof . The first part of the theorem is immediate from Lemma 8.7; the second part of thetheorem follows from (8.3) and Theorem 8.4.  

As an example, consider   A = [ 1 00 1 ] and   B = [ 1

0 ] again. The controllability Gramian is

Wc(0, T ) =

   T 

0

e−τ  0

0   e−τ 

10

1 0

e−τ  00   e−τ 

 dτ  =

(1 − e−2T )/2 0

0 0

.

We have   R(Wc(0, T )) = span{[1 0]T}   for   T >   0, so   x0   = [1 0]T is controllable, and, for everyT > 0, there is an input  u  that steers the state of the system  x(t) =  Ax(t) + Bu(t) from   x(0) =  x0

to   x(T ) = 0. With  η0 = [2/(1 − e−2T ) 0]T, we have   Wc(0, T )η0 =  x0, so such an input is given by

x0 =

10

  ⇒   u(t) = −

1 0

e−t 00   e−t

2/(1 − e−2T )

0

 = −

  2e−t

1 − e−2T  , t ∈ [0, T ).

Indeed, this   u  satisfies  eAT x0 + Lr(0, T )u =  x0 + Lc(0, T )u =  x0 +

 T 

0   e−Aτ Bu(τ ) dτ  = 0.

8.4 Controllability of Continuous-Time LTI Systems

We will now investigate the conditions under which an LTI system is reachable or controllable.

Definition 8.9   The system   (8.1), or the pair   (A, B), is said to be   reachable   if the reachable 

subspace  Rr  = Rn. Similarly, it is said to be  controllable   if the controllable subspace  Rc = Rn.

Definition 8.10   The   controllability matrix  (or   reachability matrix) of the system   (8.1)   is 

C  =

B AB   · · ·   An−1

B

∈ Rn×mn.

Lemma 8.11  The system   (8.1)  satisfies  R(Wr(0, T )) = R(C)  for every  T > 0.

Proof . To show  R(Wr(0, T ))   ⊂ R(C) for every   T >  0, suppose that   x1   ∈ R(Wr(0, T )) for

some   T >   0. Due to Lemma 8.3, we have   x1   ∈ R(Lr(0, T )), so   x1   = T 

0   eA(T −τ )Bu1(τ ) dτ   for

some   u1. Due to the Cayley-Hamilton theorem, we may write eA(T −τ ) =

 n−1k=0 β k(T  − τ )A

k forsome  β 

0(T  − τ ), . . . ,  β n−

1(T  − τ ) ∈ R. Thus

x1 =n−1k=0

Ak

B

   T 

0β k(T  − τ )u1(τ ) dτ  =

n−1k=0

Ak

Bαk(T ) =  Cα0(T )T

α1(T )T · · ·   αn−1(T )TT

,

where  αk(T ) = T 

0   β k(T  − τ )u1(τ ) dτ   ∈  Rm for  k  = 0, . . . ,  n − 1. This implies   x1  ∈ R(C) for all

T > 0, and hence establishes the desired inclusion for all  T > 0.Conversely, to show  R(C)  ⊂ R(Wr(0, T )) for every  T >  0, suppose that   x1  ∈ R(C) and that

x1   /∈ R(Wr(0, T )) for some  T > 0. It follows from   x1  ∈ R(C) that   x1  =  Cη1  for some   η1  ∈  Rmn.

Page 70: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 70/119

EE 580 — Linear Control Systems    5

On the other hand, as   Wr(0, T ) is symmetric, we have  R(Wr(0, T )) ⊕ N (Wr(0, T )) =  Rn, and

so   x1   /∈ R(Wr(0, T )) implies that   x1  =   xr +  xn   with   xr  ∈ R(Wr(0, T )),   xn  ∈ N (Wr(0, T )), andxn = 0. As   Wr(0, T )xn  = 0, we have   x

TnWr(0, T )xn  = 0. Hence it follows from

xTn Wr(0, T )xn  =  

  T 

0

xTneA(T −τ )

BBTeA

T(T −τ )xn dτ  =  

  T 

0

BTeA

T(T −τ )xn2dτ 

and the nonnegativity and continuity of   BTeA

T(T −τ )xn2 in   τ   that   x

TneA(T −τ )

B   = 0 for allτ   ∈   [0, T ]. Using the fact that (d/dt)eAt =   AeAt =  eAt

A, we obtain   xTneA(T −τ )

Ak

B  = 0 for allk ≥ 0. In particular, evaluating these equalities at  τ  = T  yields that

xnB =  xTn AB = · · · =  x

Tn A

n−1B = 0 or   x

TnC  = 0.

However,   x1 =  Cη1 for some  η1  and  xTn xr  = 0, so  x

Tn xn =  x

Tn x1  =  x

TnCη1  = 0. This contradicts   xn

being nonzero. Therefore, we must have that   x1  ∈ R(C) implies   x1  ∈ R(Wr(0, T )) for all  T > 0.This completes the proof that the equality  R(C) = R(Wr(0, T )) holds for every  T > 0.  

A consequence of the above lemma is that, if the system is reachable, the time  T   in which the

state is transferred to the desired state can be made arbitrarily small. Thus, it is theoreticallypossible to accomplish the desired state transfer instantaneously. However, in general, the fasterthe transfer, the larger the control magnitude required.

Lemma 8.12  The subspace  R(C)  is invariant under   A  (and hence under  eAt  for all   t ∈ R).

Proof . If   x   ∈ R(C), then   x   =   Cη   for some   η   ∈   Rmn, so   x   =

 n−1k=0  A

kBηk   for some   η0,

. . . ,  ηn−1 ∈ Rm. Thus, by the Cayley-Hamilton theorem, we have

Ax =

  nk=1

Ak

Bηk  =

n−1k=0

β kAk

Bηk  =  C

β 0η

T0   β 1η

T1   · · ·   β n−1η

Tn−1

T

for some  β 0, . . . ,  β n−1   ∈  R, which implies   Ax  ∈ R(C). Applying the Cayley-Hamilton theoremonce more gives that R(C) is invariant under  eAt as well regardless of  t ∈ R.  

Lemma 8.13   Rr  = Rc = R(Wr(0, T )) = R(Wc(0, T )) = R(C)  for every  T > 0.

Proof . By Lemmas 8.7 and 8.11, we have Rr   =  R(Wr(0, T )) =  R(C) for every   T >   0, so itsuffices to show that R(Wc(0, T )) = R(C) for every  T > 0. Suppose  x ∈ R(C). Using Lemma 8.12and the Cayley-Hamilton theorem, we deduce that  R(C) is invariant under  eAt for every  t ∈ R, sothat  eAT x  ∈ R(C) for all  T >  0. Then Lemma 8.11 implies eAT x  ∈ R(Wr(0, T )) for all  T >  0.Whenever  eAT 

x =  Wr(0, T )η  for some  η ∈ Rn, we have

x =  e−AT Wr(0, T )η  = e−AT 

Wr(0, T )e−ATT eA

TT η =  Wc(0, T )eA

TT η.

Thus   x  ∈ R(Wc(0, T )) for all  T >  0. This establishes that R(C)  ⊂ R(Wc(0, T )) for all  T >  0.Conversely, suppose   x ∈ R(Wc(0, T )) for some  T > 0. Then reversing the above argument showsthat   eAT 

x   ∈ R(C). Since   R(C) is invariant under   e−AT , we deduce   x   ∈ R(C). Thus we haveR(Wc(0, T )) ⊂ R(C) for all  T > 0. This completes the proof.  

Theorem 8.14   For the system   (8.1), the following are equivalent:

(a) The pair  (A, B)  is reachable;

Page 71: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 71/119

EE 580 — Linear Control Systems    6

(b) The pair  (A, B)  is controllable;

(c) The reachability Gramian   Wr(0, T ) >  0  for every  T > 0;

(d) The controllability Gramian   Wc(0, T ) >  0  for every  T > 0;

(e) The controllability matrix  C   is of full row rank; that is,  rank C  =  n;

(f) For every  T > 0, the rows of  eAtB  (as functions of   t) are linearly independent on   [0, T ];

(g) The rows of  (sI − A)−1B  are linearly independent.

Moreover, if any of these conditions holds, then for any   x0,   x1   ∈   Rn and any   T   ∈   (0, ∞), an 

input   u  that transfers the state   x   from   x(0) =  x0   to   x(T ) =  x1  is given by 

u(t) =  BTeA

T(T −t)Wr(0, T )−1

x1 − eAT 

x0

= −B

Te−ATt

Wc(0, T )−1

x0 − e−AT x1

, t ∈ [0, T ).

Proof . The equivalence of conditions (a)–(e) follows from Lemma 8.13. Since (sI − A)−1B

is the Laplace transform of   eAtB,   t   ≥   0, the equivalence of (f) and (g) follows from the factthat the Laplace transformation is a linear one-to-one correspondence (which implies that a linearcombination of the rows of (sI − A)−1

B   is zero if and only if that same linear combination of the rows of   eAtB   is zero for all   t). The forms of   u   in the last part of the theorem are fromTheorems 8.4 and 8.8. It remains to show the equivalence of (a) and (f). Suppose (a) holds but (f)does not. Then there exists a  T > 0 and an   x1 = 0 such that   x

T1 eAt

B = 0 for every  t ∈ [0, T ]. Onthe other hand, since every state is reachable within every time span, there exists a   u   such thatx1  =

 T 

0   eA(T −τ )Bu(τ ) dτ . However, as   x

T1 x1  =

 T 

0   xT1 eA(T −τ )

Bu(τ ) dτ   with   xT1 eA(T −τ )

B = 0 forall   τ   ∈   [0, T ], we must have   x1  = 0, which is a contradiction. Thus (a) implies (f). Conversely,to show (f) implies (a), suppose (a) does not hold. Then, due to the equivalence of (a) and (c),there exists a   T >   0 and an   x1   = 0 such that   x

T1 Wr(0, T ) = 0. This implies   x

T1 Wr(0, T )x1   = T 

0   B

T

eAT(T −τ )

x12

dτ  = 0, which in turn implies   x

T1 eAt

B  = 0 for all   t  ∈   [0, T ]. Thus (f) doesnot hold. That is, if (f ) holds, then (a) must hold as well. This establishes the equivalence of (a)and (f), and hence completes the proof.  

For example, consider the pairs

(A1, B1) =

1 00 1

,

10

  and (A2, B2) =

0 0 0

1 0 00 0 1

,

1 0

0 00 1

.

Their controllability matrices are

C1

 = B1

  A1

B1 = 1 1

0 0   and   C

2 = B

2  A

2B

2  A

2

2B

2 =

1 0 0 0 0 00 0 1 0 0 00 1 0 1 0 1

.

The pair (A1, B1) is not controllable because rank C1   = 1   <   2; indeed, only the states withinR(C1) = span{[1 0]T}   are controllable. On the other hand, we have rank C2   = 3, so the pair(A2, B2) is controllable. It follows from

eA2t =

1 0 0

t   1 00 0   et

, eA

T

2 t =

eA2t

T=

1   t   0

0 1 00 0   et

,

Page 72: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 72/119

EE 580 — Linear Control Systems    7

Wc(0, T ) =

   T 

0e−A2τ B2B

T2 e−A

T

2 τ  dτ  =

T    −T 2/2 0

−T 2/2   T 3/3 00 0 (1 − e−2T )/2

,

and   Wc(0, T )−1 =

4/T    6/T 2 06/T 2 12/T 3 0

0 0   −2/(e−2T 

− 1)

that, for any   x0,   x1   ∈   R3 and any   T >  0, the following input signal will drive the state of the

system  x(t) =  A2x(t) + B2u(t) from   x(0) =  x0  to   x(T ) =  x1:

u(t) = −BT2 e−A

T

2 t

Wc(0, T )−1(x0 − e−A2T x1)

=

− 4

T   +   6t

T 2  −   6

T 2 +   12t

T 3  0

0 0   2e−t

e−2T −1

x0 −

1 0 0

−T    1 00 0   e−T 

x1

, t ∈ [0, T ).

8.5 Controllability of Discrete-Time Linear Systems

In this section, we focus on the linear time-invariant state equation

x(t + 1) =  Ax(t) + Bu(t),   (8.5)

where   A ∈ Rn×n and   B ∈ R

n×m. The solution to this difference equation subject to   x(0) =  x0  canbe written as

x(t) =   At

x0 +t−1τ =0

At−τ −1

Bu(τ ) =  At

x0 + CtU(0, t)

for all  t  = 1, 2, . . . , where

Ct = B AB   · · ·   At−1

B ,

U(0, t) =

u(t − 1)T u(t − 2)T · · ·   u(0)T

T .

The definitions of reachability and controllability are analogous to their continuous-time coun-terparts. It is readily seen that the reachable subspace Rr  and controllable subspace  Rc   for thesystem (8.5) are

Rr  = {x1 ∈ Rn :   x1 ∈ R(CK ), K  = 1, 2, . . . },

Rc = {x0 ∈ Rn :   A

K x0 ∈ R(CK ), K  = 1, 2, . . . }.

The discrete-time results presented in this section can be proved analogously to their continuous-time counterparts, so their proofs are omitted. The reachability Gramian of the system (8.5) is

given by

Wr(0, K ) =K −1τ =0

AK −(τ +1)

BBT

ATK −(τ +1)

=  CK CTK .

In particular,   Wr(0, n) =  CCT, where  C  =  Cn   is the controllability matrix. Unlike the continuous-

time case, the discrete-time state transition matrix   Φ(t, t0) =   At−t0 may be singular. Therefore,unless   A   is nonsingular, the controllability Gramian for the system (8.5) is not defined. A con-sequence is that, although reachability implies controllability, controllability does not guaranteereachability in the case of discrete-time systems.

Page 73: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 73/119

EE 580 — Linear Control Systems    8

Theorem 8.15   A state   x1 ∈ Rn is reachable if and only if 

x1 ∈ R(Wr(0, K ))

 for some  K  ∈ {1, 2, . . . }. With such a  K , an input   u  that transfers the state   x   from   x(0) = 0   tox(K ) =  x1  is given by 

u(t) =  BT

AK −t−1

Tη1, t = 0, . . . , K  − 1,

where   η1 ∈ Rn is such that   Wr(0, K )η1 =  x1, or by   U(0, K )  such that 

CK U(0, K ) =  x1.

Theorem 8.16   A state   x0 ∈ Rn is controllable if and only if 

AK 

x0 ∈ R(Wr(0, K ))

 for some  K  ∈ {1, 2, . . . }. With such a  K , an input   u  that transfers the state   x   from   x(0) =   x0   tox(K ) = 0   is given by 

u(t) = −

BTA

K −t−1T

η0, t = 0, . . . , K  − 1,where   η0 ∈ R

n is such that   Wr(0, K )η0 =  AK 

x0, or by   U(0, K )  such that 

CK U(0, K ) = −AK 

x0.

Lemma 8.17  The system   (8.5)  satisfies  Rr  = R(Wr(0, K )) = R(C)  for every  K  ≥ n.

A consequence of the above lemma is that, in the case of reachable discrete-time systems, thenumber of steps  K  required to transfer the state to the desired state can be taken to be equal to  n.

Theorem 8.18   The following are equivalent:

(a) The pair  (A, B)  is reachable;

(b) The reachability Gramian   Wr(0, K ) >  0  (i.e.,  rank Wr(0, K ) = n) for  K  = n,  n + 1, . . . ;

(c) The controllability matrix  C  has full row rank (i.e.   rank C  =  n).

Moreover, if any of these conditions hold, then there exists an input that will transfer any state   x0

to any other state   x1   in some finite time  K . Such an input can be taken to be 

u(t) =  BT

ATn−t−1

Wr(0, n)−1(x1 − An

x0), t = 0, . . . , n − 1,

with  K  = n, or equivalently, to satisfy 

U(0, n) =  CT

CC

T

−1

(x1 − An

x0).

Theorem 8.19  For the system   (8.5), we have  Rr  ⊂  Rc  (i.e., reachability implies controllability).In particular, if   A  is nonsingular, then  Rc =  Rr.

References

[1] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[2] ——,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

Page 74: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 74/119

EE 580 — Linear Control Systems

IX. Observability of Linear Systems

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

9.1 Introduction

Consider the linear time-invariant system

˙x

(t) = Ax

(t) +Bu

(t),y(t) = Cx(t) + Du(t),

(9.1)

where  A ∈  Rn×n,  B ∈  R

n×m,   C ∈  Rl×n, and  D ∈  R

l×m are constant matrices. For  t,   t0  ∈  R, theoutput  y(t) of the system is given by

y(t) = CeA(t−t0)x(t0) +

   t

t0

CeA(t−τ )Bu(τ ) dτ  + Du(t).   (9.2)

We are concerned with the problem of determining the state of the system (9.1) based on the input-output data. This is an “observability” problem, which needs to be addressed before solving thecontrollability problem. Consider, for instance, the track-seeking control problem for a hard disk

drive, where the magnetic head needs to be steered from an arbitrary initial state  x(0) to a desiredterminal state x1 in a time span of  T  seconds. One strategy to approach this problem is to partitionthe time interval [0, T ) into two subintervals [0, T 1) and [T 1, T ) with some  T 1   ∈  (0, T ) and solvethe observability problem during the first subinterval and solve the controllability problem duringthe second subinterval. That is, during the first subinterval [0, T 1), one takes the measurementsy(t),   t  ∈   [0, T 1], and determine   x(0), and hence   x(T 1), based on them. Then during the secondsubinterval [T 1, T ), one steers the state from  x(T 1) to x(T ) = x1 in the time span of  T − T 1 seconds.

9.2 Unobservable States

Definition 9.1   A state  x0 ∈ Rn of the system   (9.1)   is said to be  unobservable  if 

CeAt

x0 = 0

 for all   t  ≥  0. That is, a state   x0   is unobservable if the zero-input response of the system   (9.1)   is zero for all time   t ≥ 0  whenever  x(0) = x0.

Definition 9.2   The  observability Gramian  of the system   (9.1)   is the matrix 

Wo(t0, t1) =

   t1

t0

eAT(τ −t0)CTCeA(τ −t0) dτ  ∈ R

n×n, t0 ≤ t1.

1

Page 75: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 75/119

EE 580 — Linear Control Systems    2

Clearly, we have that  Wo(t0, t1) ≥ 0 (i.e.,  Wo(t0, t1) is symmetric and nonnegative definite) for allt0,  t1 ∈ R  with  t0 ≤ t1.

Lemma 9.3   A state  x0 ∈ Rn is unobservable if and only if  x0 ∈ N (Wo(0, T ))  for all  T > 0.

Proof . If  Wo(0, T )x0  = 0 for all  T > 0, then

xT0 Wo(0, T )x0  =

   T 

0

CeAτ x02 dτ  = 0

for all  T >  0. Because CeAτ x02 is continuous in  τ , this implies that  CeAτ x0  = 0 for all  τ   ≥ 0,and hence that  x0 is unobservable. Conversely, if  x0  is unobservable, then CeAtx0  = 0 for all t ≥ 0,and so

Wo(0, T )x0  =

   T 

0eA

Tτ CTCeAτ x0 dτ  = 0

regardless of  T > 0. This implies  x0  belongs to the null space of  Wo(0, T ) for every  T > 0.  

A consequence of Lemma 9.3 is that the unobservable subspace (i.e., the subspace of all unob-servable states) is given by

Ro = {x0 ∈ Rn :  x0 ∈ N (Wo(0, T )) for all  T > 0}.

Note that the unobservable subspace  Ro   is  not  equal to the set

{x0 ∈ Rn :  x0 ∈ N (Wo(0, T )), T > 0},

which only requires that   x0   ∈ N (Wo(0, T )) for “some”   T >   0; for   x0   to be unobservable, it isrequired that  x0 ∈ N (Wo(0, T )) for “every”  T > 0.

Theorem 9.4   If   Wo(0, T )  >   0   for some   T   ∈   (0, ∞), then the unobservable subspace   Ro   =  {0}and the state  x0  at time   t = 0  is given by 

x0 =  Wo(0, T )−1    T 

0 eAT

t

CT

y(t) −   t0 Ce

A(t−τ )

Bu(τ ) dτ  − Du(t)

dt.   (9.3)

Proof . If   Wo(0, T )   >   0, then   Wo(0, T ) is nonsingular with rank Wo(0, T ) =   n, and hence N (Wo(0, T )) =  {0}. Lemma 9.3 implies that   Ro   =  {0}   in this case. It follows from (9.2) witht0  = 0 and  x(t0) = x0  that

   T 

0eA

TtCT

y(t) −

   t

0CeA(t−τ )Bu(τ ) dτ  − Du(t)

dt =

   T 

0eA

TtCTCeAtx0 dt =  Wo(0, T )x0,

which gives the desired result.  

Theorem 9.4 indicates that, if the observability Gramian  Wo(0, T )  >  0   for some  T >  0, thenthe initial state   x(0) can be uniquely determined based on the (present and future) output   y(t)

and input  u(t) over   t ∈ [0, T ].For example, consider the pair

(A, C) =

0 1 0

0 0 00 0 1

,

1 0 00 0 1

⇒   Wo(0, T ) =

   T 

0eA

Tτ CTCeAτ  dτ  =

T T 2/2 0

T 2/2   T 3/3 00 0 (e2T  − 1)/2

, T > 0.

Page 76: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 76/119

EE 580 — Linear Control Systems    3

We have that det Wo(0, T ) =   T 4(e2T  − 1)/24   >   0 for   T >   0, so   Wo(0, T ) is nonsingular, orequivalently (because  Wo(0, T ) ≥  0),  Wo(0, T )  >  0   for  T >   0. Based on the zero-input responsey(t) of the system (9.1) over  t ∈ [0, T ] for any T > 0, one can uniquely determine the state  x(0) as

x(0) = Wo(0, T )−1    T 

0

eATtCTy(t) dt =

4/T    −6/T 2 0−6/T 2 12/T 3 0

0 0 2/(e2T  − 1)

 

  T 

0

1 0t   0

0   et

y(t) dt.

If the zero-input response is

y(t) = CeAtx(0) =

1 0 00 0 1

1   t   00 1 00 0   et

−1

12

 =

t − 12et

for  t ∈ R, then indeed we obtain

x(0) =

4/T    −6/T 2 0

−6/T 2 12/T 3 0

0 0 2/(e

2T 

− 1)

T 2/2 − T 

T 3/3 − T 2/2

e

2T 

− 1

 =

−11

2

using the above integral equation.

9.3 Unconstructible States

Definition 9.5   A state  x1 ∈ Rn of the system   (9.1)   is said to be  unconstructible  if 

CeAtx1 = 0

 for all   t ≤ 0. That is, a state  x1   is unconstructible if the zero-input response of the system   (9.1)  is zero for all time   t ≤ 0  whenever  x(0) = x1.

Definition 9.6   The   constructibility Gramian  of the system   (9.1)  is the matrix 

Wcn(t0, t1) =

   t1

t0

eAT(τ −t1)CTCeA(τ −t1) dτ  ∈ R

n×n, t0 ≤ t1.

Clearly, we have that   Wcn(t0, t1)  ≥  0   for all   t0,   t1  ∈  R  with   t0  ≤  t1. The observability Gramianand the constructibility Gramian are related via the identity

Wo(t0, t1) = eAT(t1−t0)Wcn(t0, t1)eA(t1−t0).

Lemma 9.7   A state  x1 ∈ Rn is unconstructible if and only if  x1 ∈ N (Wcn(0, T ))  for all  T > 0.

Proof . The proof is analogous to that of Lemma 9.3.  

By Lemma 9.7 we see that the unconstructible subspace is given by

Rcn  = {x1 ∈ Rn :  x1 ∈ N (Wcn(0, T )) for all  T > 0}.

Theorem 9.8   If  Wcn(0, T ) >  0  for some  T  ∈ (0, ∞), then the unconstructible subspace  Rcn  = {0}and the state  x1  at time   t =  T   is given by 

x1 =  Wcn(0, T )−1   T 

0eA

T(t−T )CT

y(t) −

   t

CeA(t−τ )Bu(τ ) dτ  − Du(t)

dt.   (9.4)

Page 77: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 77/119

EE 580 — Linear Control Systems    4

Proof . If  Wcn(0, T )  >  0, then   Wcn(0, T ) is nonsingular with rank Wcn(0, T ) =  n, and hence N (Wcn(0, T )) = {0}. Lemma 9.7 implies that Rcn  =  {0}   in this case. It follows from (9.2) witht0  =  T   and  x(t0) = x1   that

   T 

0

eAT(t−T )CT

y(t) −

   t

CeA(t−τ )Bu(τ ) dτ  − Du(t)

dt =

   T 

0

eAT(t−T )CTCeA(t−T )x1 dt

= Wcn(0, T )x1,

which gives the desired result.  

Theorem 9.8 says that, if the constructibility Gramian  Wcn(0, T ) >  0  for some  T > 0, then thestate  x(T ) at time  T  can be uniquely determined based on the (past and present) output  y(t) andinput  u(t) over  t ∈ [0, T ]. To illustrate this result, consider the quadruple

(A, B, C, D) =

0 1 0

0 0 00 0 1

,

0

10

,

1 0 00 0 1

,

00

.

In this example, the constructibility Gramian is

Wcn(0, T ) =

   T 

0eA

T(τ −T )CTCeA(τ −T ) dτ  =

T    −T 2/2 0

−T 2/2   T 3/3 00 0 (1 − e−2T )/2

, T > 0.

We have that det Wcn(0, T ) = T 4(1 − e−2T )/24 >  0 for  T > 0, so  Wcn(0, T ) is nonsingular as wellas  Wcn(0, T ) ≥ 0, or equivalently,  Wcn(0, T ) >  0  for  T > 0. Suppose

x(0) = 0;   u(t) = 1, t ≥ 0;

y(t) =    t

0

CeA(t−τ )Bu(τ ) dτ   =    t

0

1 0 0

0 0 1

1   t − τ    00 1 0

0 0   et−τ 

01

0

 dτ  =

t2/2

0 , t ≥ 0.

Then the state  x(T ) is uniquely determined (without knowing  x(0) = 0) as

x(T ) = Wcn(0, T )−1   T 

0eA

T(t−T )CT

y(t) −

   t

CeA(t−τ )Bu(τ ) dτ 

dt

=

4/T    6/T 2 0

6/T 2 12/T 3 00 0 2/(1 − e−2T )

   T 

0

1 0

t − T    00   et−T 

t2/2

0

   t

t − τ 

0

 dτ 

dt

=

T 2/2

0

.

Indeed,  x(T ) = T 

0   eA(T −τ )Bu(τ ) dτ   = [T 2/2   T    0]T.

9.4 Observability of Continuous-Time LTI Systems

Definition 9.9  The system   (9.1)  is said to be   observable   if there exists  T  ∈ (0, ∞)  such that  x(0)can be uniquely determined based on  y(t),   u(t),   t ∈ [0, T ]. Similarly, it is said to be   constructibleif there exists  T  ∈ (0, ∞)  such that  x(T )  can be uniquely determined based on  y(t),  u(t),  t ∈ [0, T ].

Page 78: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 78/119

EE 580 — Linear Control Systems    5

Definition 9.10   The  observability matrix  of the system   (9.1)   is 

O =

C

CA...

CAn−1

∈ Rln×n.

Lemma 9.11  The system   (9.1)  satisfies  N (Wo(0, T )) =  N (O)  for every  T > 0.

Proof . Suppose  x0 ∈ N (Wo(0, T )) for some  T > 0, so that  Wo(0, T )x0  = 0. Then

xT0 Wo(0, T )x0  =

   T 

0CeAτ x02 dτ  = 0

for some   T >   0. Because   CeAτ x02 is continuous in   τ , this implies that   CeAτ x0   = 0 for allτ  ∈ [0, T ]. Taking the derivatives of  CeAτ x0  with respect to  τ , we have that

CeAτ 

x0  =  CAeAτ 

x0 = · · · = CA

n−1

eAτ 

x0  = 0

for all  τ  ∈ [0, T ]. In particular, these equalities at  τ  = 0 give

Cx0  =  CAx0 = · · · = CAn−1x0  = 0.

Thus Ox0 = 0, or x0 ∈ N (O). This shows that  N (Wo(0, T )) ⊂ N (O) for each T > 0. Conversely,suppose   x0   ∈ N (O). Then we have   CAkx0   = 0 for   k   = 0, . . . ,   n − 1, which, along with theCayley-Hamilton theorem, implies  CAkx0  = 0 for all  k  = 0, 1, . . . . Since  eAτ  =

 ∞

k=0(tk/k!)Ak,we have   CeAτ x0   = 0 for all   τ   ∈   R, which implies   Wo(0, T )x0   = 0 regardless of   T >   0. Thus

 N (O) ⊂ N (Wo(0, T )) for every  T > 0.  

Lemma 9.12  The system   (9.1)  satisfies  N (Wcn(0, T )) =  N (O)  for every  T > 0.

Proof . Suppose  Wcn(0, T )x1  = 0 for some  T >   0. Then  xT1 Wcn(0, T )x1  = 0, or equivalently 

0   CeA(τ −T )x12 dτ   = 0. Since  CeA(τ −T )x12 is continuous in  τ , this implies   CeA(τ −T )x1  = 0for all  τ  ∈ [0, T ]. Taking the derivatives of both sides of this equation with respect to  τ   we obtain

CeA(τ −T )x1 = CAeA(τ −T )x1 = · · · = CAn−1eA(τ −T )x1 = 0;

in particular, putting   τ   =  T   gives   Cx1  =  CAx1   =   · · ·  =   CAn−1x1   = 0. Thus  Ox1  = 0, whichimplies  N (Wcn(0, T ))   ⊂ N (O) for every   T >   0. Conversely, suppose  Ox1   = 0. Then, by theCayley-Hamilton theorem, we have  CAkx1 = 0 for all  k = 0, 1, ..., so with any  T > 0 we have

   0−T 

eATtCTCeAtx1 dt =   T 0

eAT(τ −T )CTCeA(τ −T ) dtx1 = Wcn(0, T )x1  = 0.

Thus, Wcn(0, T )x1  = 0 for every  T > 0.  

Lemma 9.13   Ro  =  Rcn  =  N (Wo(0, T )) =  N (Wcn(0, T )) = N (O)  for every  T > 0.

Proof . By Lemmas 9.3 and 9.11 we have Ro  =  N (Wo(0, T )) =  N (O) for all  T > 0. Similarly,by Lemmas 9.7 and 9.12 we have  Rcn  =  N (Wcn(0, T )) =  N (O) for all  T > 0.  

Page 79: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 79/119

EE 580 — Linear Control Systems    6

Lemma 9.14  The system   (9.1)  is observable if and only if  Ro = {0}. Similarly, it is constructible if and only if  Rcn  = {0}.

Proof . Suppose Ro   =   {0}. Then by Lemma 9.13 we have   Wcn(0, T )   >   0   for every   T >   0,and hence by Theorem 9.4 the system is observable. Now, suppose  Ro = {0}. Then, since  Ro   is asubspace of  Rn, there exists a nonzero  x0  ∈  R

n such that   CeAtx0  = 0 for all   t  ≥   0. Choose anyx0 ∈ Rn. Then the zero-input response  y(t),  t ≥ 0, for  x(0) = x0  will be exactly the same as thatfor  x(0) = x0 +  x0. Thus the system is not observable. This proves that the system is observableif and only if  Ro  = {0}. The fact that the system is constructible if and only if  Rcn  = {0}  can beproved in a similar manner using Lemma 9.13 and Theorem 9.8.  

A consequence of Lemmas 9.13 and 9.14, and Theorem 9.4 is that, if the system is observable,the time span   T  over which the input and output are observed and collected to determine theinitial state can be made arbitrarily small. Thus, it is theoretically possible to accomplish stateobservation instantaneously. However, a very small  T  may lead to a nearly singular observabilityGramian, which can lead to numerical difficulties. Moreover, Lemmas 9.13 and 9.14 suggest thatobservability and constructibility depend only on the matrices  A  and  C. Therefore, one can speakof the observability and constructibility of the pair (A, C) to mean those of the system (9.1).

Theorem 9.15   For the system   (9.1), the following are equivalent:

(a) The pair  (A, C)   is observable;

(b) The pair  (A, C)  is constructible;

(c) The observability Gramian  Wo(0, T ) >  0  for every  T > 0;

(d) The constructibility Gramian   Wcn(0, T ) >  0  for every  T > 0;

(e) The observability matrix  O   is of full column rank; that is,  rankO = n;

(f) For every  T > 0, the columns of  CeAt (as functions of  t) are linearly independent on   [0, T ];

(g) The columns of  C(sI − A)−1 are linearly independent.

Moreover, if condition (c) holds, then the state   x0   at   t = 0  is given by   (9.3)   for each  T > 0, and,if condition (d) holds, then the state   x1  at   t =  T   is given by   (9.4)   for each  T > 0.

Proof . The equivalence of (a)–(e) follows from Lemmas 9.13 and 9.14. Since  C(sI − A)−1 isthe Laplace transform of  CeAt,  t ≥ 0, the equivalence of (f) and (g) follows from the fact that theLaplace transformation is a linear one-to-one correspondence. The form of  x0   in the last part of the theorem is due to Theorem 9.4. Thus, to complete the proof, it suffices to show that (c) and (f)are equivalent. Suppose (c) holds but (f) does not. Then there exists a T > 0 and an  x0 = 0 suchthat  CeAtx0  = 0 for all t ∈ [0, T ]. On the other hand, since Wo(0, T ) >  0  with such a  T , the initial

state  x(0) = x0  is uniquely recovered based on the zero-input response  y(t),   t ∈ [0, T ], via

x0 =  Wo(0, T )−1   T 

0eA

TtCTy(t) dt =  Wo(0, T )−1

   T 

0eA

TtCTCeAtx0 dt,

which implies   x0   = 0, which contradicts the supposition that   x0   = 0. This proves that (c) im-plies (f). To prove the converse, suppose (f) holds but not (c). Then, for some T > 0 and  x0 = 0,

we have Wo(0, T )x0  = 0, and hence xT0 Wo(0, T )x0  =

 T 

0   CeAτ x02 dτ , which implies CeAτ x0 = 0for all  τ  ∈ [0, T ]. This contradicts (f), and hence proves that (f) implies (c).  

Page 80: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 80/119

EE 580 — Linear Control Systems    7

As an example, let us revisit

(A, C) =

0 1 0

0 0 00 0 1

,

1 0 00 0 1

.

This pair gives the observability matrix

O =

C

CA

CA2

 =

1 0 00 0 10 1 00 0 10 0 00 0 1

,

whose rank is 3. Thus the pair (A, C) is observable (and constructible) as seen at the end of theprevious sections.

9.5 Observability of Discrete-Time Linear Systems

Consider the state equation

x(t + 1) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t)  (9.5)

where  A ∈ Rn×n,  B ∈ R

n×m,  C ∈ Rl×n, and  D ∈ R

l×m are constant matrices. The output  y(t) of this system is given by

y(t) = CAt−t0x(t0) +t−1

s=t0

CAt−s−1Bu(s) + Du(t)

For all  t,  t0 ∈ Z  with  t ≥ t0. Thus we may write

Y(0, t) = Otx(0) + MtV(0, t)

for  t > t0, where

Y(0, t) =

y(0)...

y(t − 1)

,   V(0, t) =

u(0)...

u(t − 1)

,

and

Ot =

C

CA...

CAt−1

,   Mt  =

D 0   · · ·   0

CB D   · · ·   0...

  ...  . . .

  ...CAt−2B CAt−3B   · · ·   D

.

Definition 9.16   The system   (9.5), or the pair   (A, C), is said to be   observable   in discrete time if there exists  K   ∈ {1, 2, . . . }   such that   x(0)  can be uniquely determined based on   y(t),   u(t),   t  ∈{0, . . . , K  − 1}.

Page 81: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 81/119

EE 580 — Linear Control Systems    8

The observability Gramian of the system (9.5) is given by

Wo(0, K ) =

K −1τ =0

AT

τ CTCAτ  = O

TK OK ;

in particular,  Wo(0, n) = O

T

O, where O = On   is the observability matrix.

Theorem 9.17  The system   (9.5)   is observable if and only if 

OTK OK  > 0   (i.e.,   rank

O

TK OK 

 =  n  )

 for some  K  ≤ n. In this case, the state  x0  at time   t = 0  is given by 

x0 =O

TK OK 

−1O

TK 

Y(0, K ) − MK V(0, K )

.

A consequence of the above theorem is that, in the case of observable discrete-time systems, thenumber of steps  K  required to determine the initial state can be taken to be at most equal to  n.

Theorem 9.18   The following are equivalent:

(a) The pair  (A, C)  is observable in discrete time;

(b) The observability Gramian  OTK OK  > 0  (i.e.,  rank

O

TK OK 

 =  n) for  K  = n,  n + 1, . . . ;

(c) The observability matrix  O  has full column rank (i.e.   rankO = n).

Moreover, if any of these conditions holds, then the state  x0  at   t = 0  is given by 

x0 = (OTO)−1OT

Y(0, n) − MnV(0, n)

.

References

[1] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[2] ——,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

Page 82: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 82/119

EE 580 — Linear Control Systems

X. Structural Properties of Linear Systems (Part 1)

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

10.1 Introduction

Recall that a linear map, or its matrix representation, between finite-dimensional vector spaces aredecomposed into Jordan blocks, which completely characterize the stability properties of the linear

map. Similar things can be done to linear dynamical systems, or their state-space representations.Consider the linear time-invariant system

x(t) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t),  (10.1)

where  A  ∈  Rn×n,   B ∈  R

n×m,  C  ∈  Rl×n, and  D  ∈  R

l×m are given matrices. If the system (10.1)is not controllable (i.e., not reachable), then it is possible to separate the controllable part of thesystem from the uncontrollable part. Similarly, if the system is not observable, then it is possibleto separate the observable part from the unobservable part.

10.2 Standard Form for Uncontrollable Systems

Lemma 10.1   Let   C   be the controllability matrix of the pair   (A, B). If   rankC   =  nr, then there exists a nonsingular matrix  Q ∈ R

n×n such that 

Q−1AQ =

A1   A12

0 A2

  and    Q−1B =

B1

0

,   (10.2)

where  A1 ∈ Rnr×nr ,  B1 ∈ R

nr×m, and the pair  (A1, B1)  is controllable.

Proof . Let   {v1, . . . , vnr}   be a basis for the controllable subspace   R(C). Since   R(C) is in-

variant under   A, there exists a unique matrix   A1   ∈   Rnr×nr , whose columns are the coordinaterepresentations of  Av1, . . . ,  Avnr

 with respect to  {v1, . . . , vnr}, such that

A v1   · · ·   vnr = v1   · · ·   vnrA1.

On the other hand, if an  x ∈ Rn is such that  x =  Bu  for some  u ∈ Rm, then

x = [B AB   · · ·   An−1B][uT 0   · · ·   0]T,

so  R(B)  ⊂ R(C). Thus there exists a unique  B1   ∈   Rnr×m , whose columns are the coordinate

representations of the columns of  B with respect to  {v1, . . . , vnr}, such that

B =

v1   · · ·   vnr

B1.

1

Page 83: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 83/119

EE 580 — Linear Control Systems    2

Now, choosing any linearly independent vectors  qnr+1, . . . ,  qn  such that

Q =

v1   · · ·   vnr  qnr+1   · · ·   qn

is nonsingular, we have

A v1   · · ·   vnr =  Q A1

0   and   A qnr+1   · · ·   qn =  Q A12

A2 for some  A12 ∈ Rnr×(n−nr) and  A2 ∈ R

(n−nr)×(n−nr); similarly

B =  Q

B1

0

.

This leads to (10.2). It remains to show that (A1, B1) is controllable. The controllability matrix C  of the pair

Q−1AQ, Q−1B

 is given by

 C =

B1   A1B1   · · ·   An−1

1   B1

0 0   · · ·   0

 =  Q−1

C.   (10.3)

Here, the first equation in (10.3) implies that the rank of  C  is equal to the rank of B1   A1B1   · · ·   An−1

1   B1

,

where  n  ≥  nr. By the Cayley-Hamilton theorem, however, the columns of this matrix are linearcombinations of the columns of the controllability matrix of (A1, B1) given by

C1 =

B1   A1B1   · · ·   Anr−11   B1

∈ R

nr×mnr ,

whose rank is at most  nr. On the other hand, the second equation in (10.3) says that rank C   =rankC =  nr. Therefore, we have rank C1 =  nr, and hence the pair (A1, B1) is controllable.  

If  Q   is as in Lemma 10.1, then the change of state variables given by  Qx(t) =  x(t) results inan equivalent system of the form ˙x1(t)

˙x2(t)

 =A1   A12

0 A2

x1(t)x2(t)

+B1

0

u(t),

y(t) =

C1   C2

x1(t)x2(t)

+ Du(t),

(10.4)

where  x(t) = [x1(t)T x2(t)]T with  x1(t)   ∈   Rnr and  x2(t)   ∈   R

n−nr , and where (A1, B1) is con-trollable. This form of state-space representations is called the  standard form for uncontrollable systems . The eigenvalues of  A1  (resp.   A2) are called  controllable eigenvalues  (resp.   uncontrollable eigenvalues ) of the system (10.1). Recall that the inverse Laplace transform of (sI − A)−1 yields

e

At

=

 p

i=1

mi−1

k=0

Aikt

k

e

λit

with

Aik  =  1

k!(mi − 1 − k)!  lims→λi

(s − λi)

mi(sI − A)−1mi−1−k

,

where  λ1, . . . ,  λ p   are distinct eigenvalues of   A  and  m1, . . . ,  m p   are their algebraic multiplicities,and where the terms  tkeλit are called the  modes  of the system (10.1). In particular, the modes of the system corresponding to controllable eigenvalues (resp. uncontrollable eigenvalues) are calledcontrollable modes   (resp.  uncontrollable modes ) of the system.

Page 84: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 84/119

EE 580 — Linear Control Systems    3

Lemma 10.2  The input-output description of the system  (10.1) is determined solely by its control-lable part; that is, if   (10.1)  and   (10.4)  are equivalent, then the transfer function matrix of   (10.1)is given by 

 H(s) = C(sI − A)−1B + D =  C1(sI − A1)−1B1 + D.

Proof . If  Q  is as in Lemma 10.1, then we have

eQ−1AQτ  = Q−1eAτ Q   and   e

A1   A12

0 A2

τ 

=

eA1τ  ∗

0   eA2τ 

,

where the symbol  ∗  denotes the block that we do not care. Hence

CeAτ B =

CQ

Q−1eAτ Q

Q−1B

 =

C1   C2

eA1τ  ∗0   eA2τ 

B1

0

 =  C1eA1τ B1,

from which the result follows.  

As a simple example, consider   A   = [ 1 10 0 ],   B   = [ 1

0 ], and   C   = [ 1 2 ]. Then   C   = [ 1 10 0 ], so the

reachable subspace  R(C) = span{v1}  with  v1 = [ 10 ]. Choose  v2 = [ 0

1 ], and let  Q = [ v1   v2 ] = [ 1 00 1 ]

to obtain   Q−1AQ = A1   A12

0 A2  = [ 1 1

0 0 ],   Q−1B  = B1

0  = [ 1

0 ], and   CQ = [ C1   C2 ] = [ 1 2 ]. Thus,

among the two eigenvalues (namely, 1 and 0) of the system, the eigenvalue 1 is controllable and theeigenvalue 0 is uncontrollable. Indeed, the transfer function of the system is

H(s) = C(sI − A)−1B =

  1s−1

2s−1s(s−1)

10

 =

  1

s − 1 = C1(sI − A1)−1B1.

Since rankO   = rank[ 1 21 1 ] = 2, both eigenvalues are observable. However, the uncontrollable

eigenvalue 0 is never excited by the input, and hence does not appear as a system pole.

10.3 Standard Form for Unobservable Systems

Lemma 10.3   Let  O   be the observability matrix of the pair   (A, C). If   rankO   =  no, then there exists a nonsingular matrix  Q ∈ R

n×n such that 

Q−1AQ =

A1   0

A21   A2

  and    CQ =

C1   0

,   (10.5)

where  A1 ∈ Rno×no,  C1 ∈ Rl×n0, and the pair  (A1, C1)   is observable.

Proof . Let   {vno+1, . . . , vn}   be a basis for the unobservable subspace  N (O). Since  N (O) isinvariant under   A, there exists a unique matrix   A2   ∈   R

(n−no)×(n−no), whose columns are thecoordinate representations of  Avno+1, . . . ,  Avn  with respect to {vno+1, . . . , vn}, such that

A vno+1   · · ·   vn = vno+1   · · ·   vn

A2.

On the other hand, if  Ox = 0, then  Cx = 0, so  N (O) ⊂ N (C). Thus we have

C

vno+1   · · ·   vn

 =  0.

Choose q1, . . . ,  qno be linearly independent vectors such that

Q =

q1   . . .   qno  vno+1   · · ·   vn

Page 85: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 85/119

EE 580 — Linear Control Systems    4

is nonsingular. Then we have

A

q1   · · ·   qno

 =  Q

A1

A21

  and   A

vno+1   · · ·   vn

 =  Q

 0

A2

for some  A1  ∈  R

no×no and  A21  ∈  R(n−no)×no . Similarly,  CQ = [C1  0] for some  C1  ∈  R

l×no. This

gives (10.5). Moreover, the observability matrix O  of the pair Q−1AQ, CQ  is given by

 O =

CT

1   AT1 CT

1   · · ·

An−11

TCT

1

0 0   · · ·   0

T

=  OQ.

Here, the first equation, along with the Cayley-Hamilton theorem, implies that rank O   is equal tothe rank of the observability matrix

O1  =

CT1   AT

1 CT1   · · ·

An−1

1

TCT

1

T∈ R

lno×no

of (A

1,C

1), which is at most   no; however, the second equation gives rank O   = rankO

  =   no.Therefore, we have rankO1 =  no, and hence (A1, C1) is observable.  

If  Q   is as in Lemma 10.3, then the change of state variables given by  Qx(t) =  x(t) results inan equivalent system of the form

˙x1(t)˙x2(t)

 =

A1   0

A21   A2

x1(t)x2(t)

+

B1

B2

u(t),

y(t) =

C1   0 x1(t)

x2(t)

+ Du(t),

(10.6)

where x(t) = [x1(t)T x2(t)]T with x1(t) ∈ Rno and x2(t) ∈ R

n−no, and where (A1, C1) is observable.This form of state-space representations is called the  standard form for unobservable systems . The

eigenvalues of  A1  (resp.   A2) and the corresponding modes are called  observable eigenvalues  (resp.unobservable eigenvalues ) and  observable modes   (resp.  unobservable modes ) of the system (10.1).

Lemma 10.4  The input-output description of the system   (10.1) is determined solely by its observ-able part. That is, if   (10.1)  and   (10.6)  are equivalent, then the transfer function matrix of   (10.1)is given by   H(s) = C(sI − A)−1B + D =  C1(sI − A1)−1B1 + D.

Proof . The proof parallels that of Lemma 10.2.  

For example, suppose  A  = [ 1 10 0 ],   B  = [ 1

1 ], and   C = [ 1 1 ]. Since  O  = [ 1 11 1 ], the unobservable

subspace  N (O) = span{v2}  with   v2   =

  1−1

. Choose a   v1   = [ 10 ], and let   Q  = [ v1   v2 ] =

1 10  −1 to obtain   Q

1AQ =   A1

  0A21   A2  = [ 1 00 0 ],   Q−

1B  = B1

B2  =   2−1 , and   CQ = [ C1   0 ] = [ 1 0 ]. Thus,among the two eigenvalues of the system, the eigenvalue 1 is observable and the eigenvalue 0 isunobservable. Indeed, the transfer function of the system is

H(s) = C(sI − A)−1B =

1 1   s+1

s(s−1)1s

 =

  2

s − 1 = C1(sI − A1)−1B1.

Since rankC  = rank[ 1 21 0 ] = 2, both eigenvalues are controllable. However, the unobservable eigen-

value 0 is never observed from the system output, and hence does not appear as a system pole.

Page 86: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 86/119

EE 580 — Linear Control Systems    5

10.4 Kalman Canonical Form of LTI Systems

The following theorem is called the  Kalman decomposition theorem . The theorem unifies the de-composition lemmas for uncontrollable and unobservable linear time-invariant systems.

Theorem 10.5   Let  C  and  O  be the controllability and observability matrices of the triple  (A, B, C).

If  rankC = nr,  rankO =  no, and  dim R(C) ∩ N (O) = nro, then there exists a nonsingular matrix Q ∈ Rn×n such that 

Q−1AQ =

A11   0 A13   0

A21   A22   A23   A24

0 0 A33   0

0 0 A43   A44

,   Q−1B =

B1

B2

0

0

,   CQ =

C1   0 C3   0

,   (10.7)

where   A11   ∈  R(nr−nro)×(nr−nro),   A22   ∈  Rnro×nro,   A33   ∈  R

(no−(nr−nro))×(no−(nr−nro)), and   A44   ∈R

((n−no)−nro)×((n−no)−nro), and such that the following hold:

(a) The pair  (Ac, Bc)  with 

Ac = A11   0A21   A22

  and    Bc = B1B2

is controllable.

(b) The pair  (Ao, Co)  with 

Ao =

A11   A13

0 A33

  and    Co =

C1   C3

is observable.

(c) The triple  (A11, B1, C1)   is controllable and observable.

Proof . Choose a basis   {v1, . . . , vnr}   for   R(C) such that   {vnr−nro+1, . . . , vnr}   is a basis forR(C) ∩ N (O). Choose vectors  vno+nro+1, . . . ,  vn  such that

{vnr−nro+1, . . . , vnr, vno+nro+1, . . . , vn   

n− no  vectors

}

is a basis for  N (O). Finally, choose  qnr+1, . . . ,  qno+nro  such that

Q =

v1   · · ·   vnr  qnr+1   · · ·   qno+nro

  vno+nro+1   · · ·   vn

is nonsingular. Then, since the first nr  columns of  Q  span R(C), (the proof of) Lemma 10.1 impliesthat  Q−1AQ,  Q−1B, and  CQ are of the following partitioned forms:

Q−1AQ =

A11

  A12

  A13

  A14

A21   A22   A23   A24

0 0 A33   A34

0 0 A43   A44

,   Q−1B =

B1

B2

0

0

,   CQ =

C1   C2   C3   C4

.   (10.8)

Let

T =

Inr−nro

  0 0 0

0 0 Inro  0

0 In0−(nr−nro)   0 0

0 0 0 I(n−no)−nro

,

Page 87: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 87/119

EE 580 — Linear Control Systems    6

where  Ik  are the  k-by-k   identity matrices. Then

QT =

v1   · · ·   vnr−nro  qnr+1   · · ·   qno+nro

  vnr−nro+1   · · ·   vnr  vno+nro+1   · · ·   vn

,

where the last n − no columns of  QT  span N (O). Thus (the proof of) Lemma 10.3, along with thefact that  T−1 = TT, implies that  Q−1AQ,  Q−1B, and  CQ are of the following partitioned forms:

Q−1AQ =  T(QT)−1A(QT)T−1 = T

F11   F12   0 0F21   F22   0 0

F31   F32   F33   F34

F41   F42   F43   F44

T−1 =

F11   0 F12   0F31   F33   F32   F34

F21   0 F22   0

F41   F43   F42   F44

,

Q−1B =  T(QT)−1B =  T

G1

G2

G3

G4

 =

G1

G3

G2

G4

,

CQ =  C(QT)T−1

H1   H2   0 0

T−1 =

H1   0 H2   0

.

Comparing these equations with those in (10.8), we conclude that (10.7) holds. The pair (Ac, Bc)

is controllable by Lemma 10.1, and the pair (Ao, Co) is observable by Lemma 10.3. It follows from

rank

Bc   AcBc   · · ·   Anr−1c   Bc

 = rank

B1   A11B1   · · ·   Anr−1

11   B1

∗ ∗ · · · ∗

 =  nr,

along with the Cayley-Hamilton theorem, that

rank

B1   A11B1   · · ·   Anr−111   B1

 = rank

B1   A11B1   · · ·   Anr−nro−1

11   B1

 =  nr − nro.

Thus the triple (A11, B1, C1) is controllable. Since (Ac, Bc, [C1  0]) is already in the standard formfor unobservable systems, we conclude that (A11, B1, C1) is observable as well.  

If  A11,  A22,  A33, and  A44  are as in Theorem 10.5, then we have

det(sI − A) = det(sI − A11)det(sI − A22)det(sI − A33)det(sI − A44).

The eigenvalues of   A11  are controllable and observable, those of   A22   are controllable and unob-servable, those of  A33  are uncontrollable and observable, and those of  A44   are uncontrollable andunobservable.

Theorem 10.6  The input-output description of the system   (10.1)  is determined solely by its con-trollable and observable part. That is, if   Q   is as in   (10.7), then the transfer function matrix of   (10.1)  is given by  H(s) = C(sI − A)−1B + D =  C1(sI − A11)−1B1 + D.

Proof . The result is an immediate consequence of Lemmas 10.2 and 10.4.  

For example, consider

A =−1   −3   −2

1 3 31 1 0

,   B =−1 0

1 10   −1

,   C =1 1 0

2 2 1

.

The controllability and observability matrices are

C =

−1 0   −2   −1   −4   −11 1 2 0 4 20   −1 0 1 0   −1

,   O =

1 2 0 1 1 21 2 0 1 1 20 1 1 2 0 1

T

.

Page 88: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 88/119

EE 580 — Linear Control Systems    7

The controllable(or reachable) subspace  Rr   =  R(C), the unobservable subspace  Ro   =  N (O),and their intersection  Rro = R(C) ∩ N (O) are

Rr = span

01

−1

,

1−10

, Ro  = span

1−10

,   and   Rro = span

1−10

,

respectively. Then

v1  =

01

−1

,   v2 =

1−10

,   q3 =

111

  ⇒   Q =

v1   v2   q3

 =

0 1 11   −1 1

−1 0 1

.

This  Q gives

Q−1AQ =

−1 0   −1

−1 2   −70 0 1

,   Q−1B =

0 1

−1 00 0

,   CQ =

  1 0 21 0 5

.

Among the three eigenvalues of   A, the eigenvalue   −1 is both controllable and observable, theeigenvalue 2 is controllable but unobservable, and the eigenvalue 1 is observable but uncontrollable.Indeed, the transfer function matrix shows a single pole at  s = −1:

 H(s) = C(sI − A)−1B =

11

(s − (−1))−1 0 1

 =

0 1/(s + 1)0 1/(s + 1)

.

10.5 Popov-Belevitch-Hautus (PBH) Tests

We know the system (10.1) is controllable if and only if its controllability matrix has full rowrank. Similarly, the system is observable if and only if its observability matrix has full rank. Thedecomposition results presented above reveal the structural properties of linear systems, and provideadditional tests for controllability and observability. These tests facilitate further insights into thestructural properties of linear systems, and will play an important role in controller synthesisproblems.

Theorem 10.7  (PBH Eigenvector Tests)

(a) The pair  (A, B)  is controllable if and only if no eigenvector  v  of  AT satisfies  BTv = 0.

(b) The pair  (A, C)  is observable if and only if no eigenvector  v  of  A  satisfies  Cv = 0.

Proof . Suppose there exists an eigenvector   v   ∈   Cn of   AT such that   BTv   = 0. Let   λ   ∈   C

be the eigenvalue associated with   v. Then we have   vTB  = 0,   vTAB  =   λvTB   = 0,   vTA2B   =

λvTAB   = 0, . . . , an d s o   vTC   = 0, where   C   is the controllability matrix of (A, B). That is,ℜ(v)T

C  = ℑ(v)TC  = 0. Since  v = 0, we have  ℜ(v) = 0 or  ℑ(v) = 0, so rank C  < n. This proves

that the pair (A, B) being controllable implies  BTv = 0 for any eigenvector  v  of  AT. Conversely,suppose (A, B) is not controllable, so that there exists a nonsingular   Q   satisfying (10.2), whereA2  ∈  R

(n−nr)×(n−nr) with  nr  < n. Choose an eigenvalue  λ  of  AT2 , and let  v  be the corresponding

eigenvector of  AT2  . Then putting  v =

Q−1

T[0 vT]T ∈ R

n gives

ATv =

Q−1T

Q−1AQT

0v

 =

Q−1T

AT1   0

AT12   AT

2

0v

 =

Q−1T

 0λv

 =  λv

Page 89: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 89/119

EE 580 — Linear Control Systems    8

and

BTv =  BT

Q−1T

0v

 =

Q−1BT

0v

 =

BT1   0

0v

 = 0.

This shows that the pair (A, B) is controllable whenever no eigenvector  v of  AT satisfies BTv = 0,and hence that part (a) of the theorem holds. The proof of part (b) is similar.  

Corollary 10.8   (PBH Rank Tests)

(a) The pair  (A, B)  is controllable if and only if 

rank

λI − A B

 =  n

 for all eigenvalues  λ  of  A  (and hence for all  λ ∈ C).

(b) A number  λ ∈ C  is an uncontrollable eigenvalue of  A  if and only if 

rank

λI − A B

 < n.

(c) The pair  (A, C)   is observable if and only if 

rank

λI − A

C

 =  n

 for all eigenvalues  λ  of  A  (and hence for all  λ ∈ C).

(d) A number  λ ∈ C  is an unobservable eigenvalue of  A  if and only if 

rank

λI − A

C

 < n.

10.6 Kalman Decomposition of Discrete-Time Systems

For discrete-time LTI systems of the form

x(t + 1) = Ax(t) + Bu(t), t = 0,   1, . . . ;

y(t) = Cx(t) + Du(t), t = 0,   1, . . . ,

even though reachability implies controllability, controllability does not imply reachability. Simi-larly, observability implies constructibility, but the converse does not necessarily hold. Thus, wewill speak of the reachability (as well as observability) of discrete-time systems. Otherwise, allthe theorems in Sections 10.4 and 10.5, with “controllability” replaced by “reachability” and the“Laplace transform” H(s) by the “z-transform” H(z), carry over to discrete-time LTI systems

without further change.

References

[1] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[2] ——,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

Page 90: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 90/119

EE 580 — Linear Control Systems

XI. Structural Properties of Linear Systems (Part 2)

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

11.1 Introduction

In this section, we continue to study the structural properties of LTI systems. In particular,we delineate how controllability and observability are related to the system response and system

stability. We will also learn that controllability and observability are dual notions, and hence thatthey are mathematically identical notions in a sense.

11.2 Controller Canonical Form

11.2.1 SISO Case

Consider the single-input single-output linear time-invariant system

x(t) =  Ax(t) + bu(t),

y(t) =  cx(t) + du(t),(11.1)

where  A ∈ Rn×n,  b ∈ R

n×1, and  c ∈ R1×n.

Theorem 11.1   Let  p(s) =  sn + αn−1sn−1 + ⋯ + α1s + α0  be the characteristic polynomial of  A; let  Cbe the controllability matrix of  (A, b). There exists a nonsingular matrix   Q ∈ R

n×n such that 

Q−1AQ =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 1   ⋯   0⋮ ⋮ ⋱ ⋮0 0   ⋯   1

−α0   −α1   ⋯ −αn−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦,   Q−1b =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0⋮01

⎤⎥⎥⎥⎥⎥⎥⎥⎦,   and    cQ = β 0   β 1   ⋯   β n−1   (11.2)

 for some  β 0, . . . ,  β n−1 ∈ R  if and only if  rankC =  n.

Proof . Suppose rankC =  n. The system being SISO implies that  C ∈ Rn×n, and hence that  C   is

invertible. Let  q be the last row of  C

−1

. Then C

−1C

 =  I  impliesq b Ab   ⋯   An−2b An−1b = 0 0   ⋯   0 1 ;

that is,  qb =  qAb = ⋯ =  qAn−2b = 0 and  qAn−1b = 1. Define  R = [qT (qA)T ⋯ (qAn−1)T]T. It isclear that

RC =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

qb   . . .   qAn−2b qAn−1b

qAb   ⋯   qAn−1b qAnb

⋮ ⋰ ⋮ ⋮qAn−1b   ⋯   pA2n−3b qA2n−2b

⎤⎥⎥⎥⎥⎥⎥⎥⎦=

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0   ⋯   0 10   ⋯   1   ∗⋮ ⋰ ⋮ ⋮1   ⋯ ∗ ∗

⎤⎥⎥⎥⎥⎥⎥⎥⎦1

Page 91: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 91/119

EE 580 — Linear Control Systems    2

is invertible. Since  C   is invertible,  R  is invertible as well. The Cayley-Hamilton theorem gives

qAn = q  p(A) −n−1

 j=0 α jA j = −n−1

 j=0 α jqA j = − α0   α0   ⋯   αn−1R,

so we obtain

RA = ⎡⎢⎢⎢⎢⎢⎢⎢⎣

qA⋮

qAn−1

qAn

⎤⎥⎥⎥⎥⎥⎥⎥⎦= ⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 1   ⋯   0⋮ ⋮ ⋱ ⋮0 0   ⋯   1

−α0   −α1   ⋯ −αn−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦R   and   Rb = ⎡⎢⎢⎢⎢⎢⎢⎢⎣

qb⋮

qAn−2b

qAn−1b

⎤⎥⎥⎥⎥⎥⎥⎥⎦= ⎡⎢⎢⎢⎢⎢⎢⎢⎣

0⋮01

⎤⎥⎥⎥⎥⎥⎥⎥⎦.

Letting  Q = R−1 yields that  Q−1AQ,  Q−1b, and  cQ are as in (11.2).Conversely, suppose that there exists an invertible  Q such that (11.2) holds. Then

Q−1C = Q−1b   (Q−1AQ)(Q−1b)   ⋯   (Q−1AQ)n−1(Q−1b)

=

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0   ⋯   0 10   ⋯   1   ∗⋮ ⋰ ⋮ ⋮

1   ⋯ ∗ ∗

⎤⎥⎥⎥⎥⎥⎥⎥⎦

.

Clearly, we have rankQ−1C = n. Since  Q−1 is invertible, we conclude that rankC =  n.   ◻

Theorem 11.2  If there exists a  Q  ∈ Rn×n such that   (11.2)  holds, then the transfer function of the 

system   (11.1)  is given by 

h(s) =  β n−1sn−1 + ⋯ + β 1s + β 0

sn + αn−1sn−1 + ⋯ + α1s + α0

+ d.

Proof . We have

(sI − Q−1AQ)⎡⎢⎢⎢⎢⎢⎢⎢⎣1s⋮

sn−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦= ⎡⎢⎢⎢⎢⎢⎢⎢⎣

0⋮0

 p(s)⎤⎥⎥⎥⎥⎥⎥⎥⎦

,   and hence  1

 p(s) ⎡⎢⎢⎢⎢⎢⎢⎢⎣1s⋮

sn−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦= (sI − Q−1AQ)−1Q−1b,

where  p(s) = sn + αn−1sn−1 + ⋯ + α1s + α0. Now the result is immediate from

c(sI − A)−1b = cQ(sI − Q−1AQ)−1Q−1b =  1

 p(s) β 0   β 1   ⋯   β n−1 1   s   ⋯   sn−1T.   ◻

11.2.2 MIMO Case

Consider the multiple-input multiple-output linear time-invariant system

x(t) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t),(11.3)

where  A ∈ Rn×n,  B ∈ R

n×m, and  C ∈ Rl×n.

If   C   is the controllability matrix of the pair (A, B)  and if   b1, . . . ,   bm   ∈  Rn are the columnsof  B such that  B = [b1 ⋯  bm], then we may write

C = b1   ⋯   bm   Ab1   ⋯   Abm   ⋯   An−1b1   ⋯   An−1bm .   (11.4)

Page 92: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 92/119

EE 580 — Linear Control Systems    3

Suppose rankC   =  nr. Starting from the left and moving to the right, select the first   nr   linearlyindependent columns of (11.4). Reorder these columns and obtain

M = b1   Ab1   ⋯   Aµ1−1b1   ⋯   bm   Abm   ⋯   Aµm−1bm .   (11.5)

The integer  µi   is the number of linearly independent columns in  M  that are associated with  bi; set

µi =  0 if no column of  M  involves  bi. The integers  µ1, . . . ,  µm  are called the  controllability indices of the pair (A, B). Clearly, we have ∑m

i=1 µi =  rankC.If the  m  columns  b1, . . . ,  bm  ∈  Rn of  B  are not linearly independent (i.e., if rank B =  r  <  m),

then some of the controllability indices are zero. In this case, there exists a matrix   Br   ∈   Rn×r

with linearly independent columns   v1, . . . ,   vr   ∈   Rn such that   b j   =   p1 jv1 + ⋯ +  prjvr   for some p1 j, . . . ,   prj   ∈  R   and for all   j   =   1, . . . ,   m. If we let   Pr   = ( pij)   ∈  Rr×m, then   B  =   BrPr. Sincerank B =  rank Br, we must have rank Pr =  r; that is, the rows of  Pr  must be linearly independent.Now choose any Ps ∈ R

(m−r)×m such that P = [PTr   PT

s ]T is nonsingular. Then we have B = [Br  0]P.Thus, whenever rank B =  r  <  m, there exist a matrix  Br ∈ R

n×r with linearly independent columnsand a nonsingular matrix  K ∈ R

m×m such that

BK =

 Br   0

,

which implies  Bu = [Br   0]K−1u. That is, only the first r  entries of the transformed input  K−1u

affect the system state. Moreover, it follows from

rank Br   ABr   ⋯   An−1Br = rank⎛⎜⎝B AB   ⋯   An−1B

⎡⎢⎢⎢⎢⎢⎣K 0

⋱0 K

⎤⎥⎥⎥⎥⎥⎦⎞⎟⎠

that (A, Br)   is controllable if and only if  (A, B)  is controllable. Therefore, without loss of gener-ality, we assume that the  m  columns of  B are linearly independent.

Theorem 11.3   Let   C   be the controllability matrix of  (A, B). If   rank B   =   m   and   rankC   =   n,and if  µ1, . . . ,  µm   are the controllability indices of 

 (A, B

), then there exists a nonsingular matrix 

Q ∈ Rn×n such that 

Q−1AQ =

⎡⎢⎢⎢⎢⎢⎣A11   ⋯   A1m

⋮ ⋮Am1   ⋯   Amm

⎤⎥⎥⎥⎥⎥⎦(11.6a)

with 

Aii =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 1   ⋯   0⋮ ⋮ ⋱ ⋮0 0   ⋯   1∗ ∗ ⋯ ∗

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

µi×µi ;   Aij =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0   ⋯   0⋮ ⋱ ⋮0   ⋯   0∗ ⋯ ∗

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

µi×µj , i ≠  j; (11.6b)

and 

Q−1B = ⎡⎢⎢⎢⎢⎢⎣B11   ⋯   B1m

⋮ ⋮Bm1   ⋯   Bmm

⎤⎥⎥⎥⎥⎥⎦(11.6c)

with 

Bii =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0⋮01

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

µi×1;   Bij  =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0⋮0∗

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

µi×1, i <  j;   Bij =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0⋮00

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

µi×1, i > j.   (11.6d)

(The matrix   CQ  does not have any particular structure.)

Page 93: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 93/119

EE 580 — Linear Control Systems    4

Sketch of Proof . Define  σk  = ∑ki=1 µi,  k =  1, . . . ,  m. Let  qk  be the  σk-th row of  M−1, where  M

is as in (11.5); that is,

M−1 = ∗ ⋯ ∗   qT1   ⋯ ∗ ⋯ ∗   qT

mT.

Use  q1, . . . ,  qm  to form

R = qT1   (q1A)T ⋯   (q1Aµ1−1)T ⋯   qTm   (qmA)T ⋯   (qmAµm−1)TT .

Then it can be shown, proceeding as in the proof of Theorem 11.1, that  R  is nonsingular and thatQ =  R−1 is the desired similarity transformation.   ◻

Suppose that  Q   is as in Theorem 11.3; suppose rankC =  n  and rank B =  m. If  µ1, . . . ,  µm  arethe controllability indices of  (A, B)  with ∑m

i=1 µi =  n, let

Ac =

⎡⎢⎢⎢⎢⎢⎣A1   0

⋱0   Am

⎤⎥⎥⎥⎥⎥⎦∈ R

n×n with   Ai =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0 1   ⋯   0⋮ ⋮ ⋱ ⋮0 0   ⋯   10 0   ⋯   0

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

µi×µi , i = 1, . . . , m; (11.7a)

Bc = ⎡⎢⎢⎢⎢⎢⎣B1   0

⋱0   Bm

⎤⎥⎥⎥⎥⎥⎦∈ R

n×m with   Bi =⎡⎢⎢⎢⎢⎢⎢⎢⎣

0⋯01

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

µi×1 i = 1, . . . , m .   (11.7b)

Then we haveQ−1AQ =  Ac +  BcAq   and   Q−1B =  BcBq   (11.7c)

for some   Aq   ∈  Rm×n and   Bq   ∈  Rm×m. If   q1, . . . ,   qm   and   Q   are as in the sketch of the proof of Theorem 11.3, then it can be shown that

Aq =

⎡⎢⎢⎢⎢⎢⎣

q1Aµ1

⋮qmAµm

⎤⎥⎥⎥⎥⎥⎦

Q   and   Bq  =

⎡⎢⎢⎢⎢⎢⎣

q1Aµ1−1

⋮qmAµm−1

⎤⎥⎥⎥⎥⎥⎦

B.

Furthermore, the matrix  Bq   is in fact an upper triangular matrix with ones on the diagonal, andhence it is nonsingular. The following is the controllable version of the, so-called,  structure theorem .

Theorem 11.4   If  Q   is as in Theorem 11.3, and if  Aq   and  Bq  are as in   (11.7), then the transfer  function matrix of the system   (11.3)  is given by   H(s) = N(s)D(s)−1 + D, where 

D(s) =  B−1q Λ(s) − AqS(s),   N(s) = CQS(s),

Λ

(s

) =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

sµ1

sµ2

⋱sµm

⎤⎥⎥⎥⎥⎥⎥⎥⎦

,   S

(s

) = diag

⎧⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎩

⎡⎢⎢⎢⎢⎢⎢⎢⎣

1s

⋮sµ1−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦

, . . . ,

⎡⎢⎢⎢⎢⎢⎢⎢⎣

1s

⋮sµm−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦

⎫⎪⎪⎪⎪⎪⎬⎪⎪⎪⎪⎪⎭

.

Proof . Since

Q−1BD(s) =  BcBqB−1q Λ(s) − AqS(s) =  BcΛ(s) −  BcAqS(s),

sI − Q−1AQS(s) = sS(s) − Ac +  BcAqS(s) = sI −  AcS(s) −  BcAqS(s)=  BcΛ(s) −  BcAqS(s),

Page 94: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 94/119

EE 580 — Linear Control Systems    5

we have sI − Q−1AQS(s) = Q−1BD(s). This gives

  H(s) = CQsI − Q−1AQ−1Q−1B + D =  CQS(s)D(s)−1 + D = N(s)D(s)−1 + D.   ◻

As an illustrative example, consider

A =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 0 3 1 2   −1−4   −3   −4 2   −1   −30 0   −2   −1   −1 1

−5   −2   −8 1   −3   −40   −1 2 2 1   −13 5 0   −4   −2 2

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦,   B =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 0   −10 1   −20 0 11 0   −50 0   −11   −2   −1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦,   C = 2 4 3   −3 1 1

1 1   −2   −1   −1 1,   D = 0 0 0

0 0 0.

Let  b1,  b2, and  b3  be the columns of  B such that  B  = [b1 b2 b3]. Then the controllability matrixof  (A, B)   is given by   C   = [b1   b2   b3   Ab1   Ab2   Ab3   ⋯  A5b1   A5b2   A5b3]. Since rankC  =  6 andrank B =  3, we can proceed to obtain the controller-form representation of  (A, B, C). Taking thefirst 6 linearly independent columns of  C  and reordering them, we obtain

M =

 b1   Ab1   A2b1   b2   b3   Ab3

=

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 0 1 0   −1   −30   −1 2 1   −2 00 0 2 0 1 11   −3 4 0   −5 30 1   −2 0   −1   −61   −2 1   −2   −1 7

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⇒   M−1 =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

−7 14 8   −6 10 75   −30   −9 15   −14   −151   −4   −1 2   −2   −2

−3 7 3   −3 4 3−3 14 5   −7 7 71   −6   −2 3   −3   −3

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

.

The controllability indices are  µ1  = 3,   µ2  = 1, and  µ3 = 2. Since  σ1  = 3,   σ2  =  4, and  σ3  = 6, takingthe 3rd, 4th, and 6th rows of  M−1, and denoting them by  q1,  q2, and  q3, respectively, we obtain

Q−1 =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

q1

q1A

q1A2

q2

q3

q3A

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

=

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1   −4   −1 2   −2   −20 0 1 0 1 00   −1 0 1 2 0

−3 7 3   −3 4 31   −6   −2 3   −3   −30 0 1 0 0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⇒   Q =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

3 0 0 0   −2   −16   −1 0 1   −3   −20 0 0 0 0 16   −3 1 1   −3 00 1 0 0 0   −1

−5   −2 1   −1 2 4

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

.

Now a controller-form representation (Ac, Bc, Cc) = (Q−1AQ, Q−1B, CQ) of  (A, B, C) is given by

Ac =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 1 0 0 0 00 0 1 0 0 0

24   −2 0 5   −10   −9−25 0 0   −2 15 0

0 0 0 0 0 1−11 2 0   −2 5 1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

,   Bc =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 0 00 0 01   −1   −50 1 00 0 00 0 1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

,   Cc =    7 4   −2 0   −5   −4−2   −1 0   −1 0 0

 .

To obtain the transfer function matrix using the structure theorem, write

Ac =  Ac +  BcAq  =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 1 0 0 0 00 0 1 0 0 00 0 0 0 0 0

0 0 0 0 0 0

0 0 0 0 0 10 0 0 0 0 0

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

+

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 0 00 0 01 0 0

0 1 0

0 0 00 0 1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎣24   −2 0 5   −10   −9

−25 0 0   −2 15 0−11 2 0   −2 5 1

⎤⎥⎥⎥⎥⎥⎦3rd, 4th, and 6th rows of  Ac

,

Page 95: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 95/119

EE 580 — Linear Control Systems    6

Bc =  BcBq  =

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

0 0 00 0 01 0 0

0 1 0

0 0 00 0 1

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎣1   −1   −50 1 00 0 1

⎤⎥⎥⎥⎥⎥⎦3rd, 4th, and 6th rows of  Bc

,

and obtain

D(s) = B−1q Λ(s) − AqS(s)

=

⎡⎢⎢⎢⎢⎢⎣1   −1   −50 1 00 0 1

⎤⎥⎥⎥⎥⎥⎦

−1

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎝

⎡⎢⎢⎢⎢⎢⎣s3 0 00   s   00 0   s2

⎤⎥⎥⎥⎥⎥⎦−

⎡⎢⎢⎢⎢⎢⎣24   −2 0 5   −10   −9

−25 0 0   −2 15 0−11 2 0   −2 5 1

⎤⎥⎥⎥⎥⎥⎦

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 0 0s   0 0

s2 0 00 1 00 0 10 0   s

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎠=

⎡⎢⎢⎢⎢⎢⎣

s3 − 8s + 56   s + 7 5s2 + 4s − 3025   s + 2   −15

−2s + 11 2   s2 − s − 5

⎤⎥⎥⎥⎥⎥⎦

,

N(s) = CQS(s) =    7 4   −2 0   −5   −4−2   −1 0   −1 0 0

 

⎡⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎣

1 0 0s   0 0

s2 0 00 1 00 0 10 0   s

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

= −2s2 + 4s + 7 0   −4s − 5−s − 2   −1 0

  .

The transfer function matrix is then given by

  H(s) = N(s)D(s)−1 + D

= 2s4 − 2s3 − 17s2 + 43s + 46 2s4 − 4s3 − 13s2 + 116s + 115 6s4 − 5s3 − 83s2 + 46

s3 + 3s2 − 30s + 61   s4 − 2s3 − s2 + 45s − 105 5s3 + 9s2 − 104s + 51s5 + s4 − 5s3 + 16s2 + 44s + 115

  .

11.3 Observer Canonical Form

11.3.1 SISO Case

Consider the single-input single-output linear time-invariant system (11.1). A dual result to The-orem 11.1 is stated as follows.

Theorem 11.5   Let  p(s) = sn + αn−1sn−1 + ⋯ + α1s + α0  be the characteristic polynomial of  A; let  Obe the observability matrix of  (A, c). There exists a nonsingular matrix  Q ∈ R

n×n such that 

Q−1AQ =⎡⎢⎢⎢⎢⎢⎢⎢⎣

0   ⋯   0   −α0

1   ⋯   0   −α1

⋮ ⋱ ⋮ ⋮0   ⋯   1   −αn−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦,   Q−1b =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

β 0β 1⋮

β n−1

⎤⎥⎥⎥⎥⎥⎥⎥⎦,   and    cQ = 0   ⋯   0 1   (11.8)

 for some  β 0, . . . ,  β n−1 ∈ R  if and only if  rankO =  n.

Sketch of Proof . The proof is analogous to that of Theorem 11.1, so it will only be sketched.Suppose rankO   =   n. Let   p   be the last column of  O−1. Then  OO

−1 =   I, so   cp  =   cAp  =  ⋯  =

Page 96: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 96/119

EE 580 — Linear Control Systems    7

cAn−2p = 0 and  cAn−1p = 1. This implies that  OQ  is invertible, where  Q  = [p Ap ⋯ An−1p], andhence that  Q  is invertible. The Cayley-Hamilton theorem then gives Anp =  −∑n−1

 j=0  α jA jp, so we

obtain that   Q−1AQ,   Q−1b, and   cQ   are of the form in (11.8). Conversely, suppose (11.8) holds.Then  OQ can be shown to have rank  n. Since  Q is invertible, we conclude that rankO =  n.   ◻

Theorem 11.6  If there exists a  Q  ∈ Rn×n such that   (11.8)  holds, then the transfer function of the 

system   (11.1)  is given by 

h(s) =  β n−1sn−1 + ⋯ + β 1s + β 0

sn + αn−1sn−1 + ⋯ + α1s + α0

+ d.

Proof . The proof is analogous to that of Theorem 11.2; the result follows from the fact that

1   s   ⋯   sn−2 sn−1 (sI − Q−1AQ) = 0 0   ⋯   0   p(s) .   ◻

11.3.2 MIMO Case

Consider the multiple-input multiple-output linear time-invariant system (11.3). If  O   is the ob-servability matrix of the pair

 (A, C

) and if  c1, . . . ,  cl ∈ R

1×n are   l  rows of  C, then we may write

O = cT1   ⋯   cT

l   (c1A)T ⋯   (clA)T ⋯   c1An−1T⋯   clA

n−1TT.   (11.9)

Suppose rankO =  no. Starting from the top and moving to the bottom, select the first  no   linearlyindependent rows of (11.9). Reorder these rows and obtain

N = cT1   (c1A)T ⋯   c1Aν 1−1T

⋯   cTl   (clA)T ⋯   clA

ν l−1TT.   (11.10)

The integer   ν i   is the number of linearly independent rows in   N   that are associated with   ci; setν i =  0 if no row of  N  involves  ci. The integers  ν 1, . . . ,  ν l  are called the  observability indices  of thepair (A, C). Clearly, we have ∑l

i=1 ν i =  rankO.If the   l   rows   c1, . . . ,   cl   of   C   are not linearly independent (i.e., if rank C  =  r  <   l), then some

of the observability indices are zero. In this case, there exist a matrix   Cr   ∈   Rr×n with linearlyindependent rows and a nonsingular matrix  L  ∈ R

l×l such that

LC = Cr

0  ,

which implies LCx = Cr

0 x. That is, only the first r  entries of the transformed output  Ly  depends

on the systems state. Moreover, (A, Cr) is observable if and only if  (A, C) is observable. Therefore,without loss of generality, we assume that the   l  rows of  C are linearly independent.

Theorem 11.7   Let  O be the observability matrix of the pair  (A, C). If  rank C =  l  and  rankO =  n,and if   ν 1, . . . ,   ν l   are the observability indices of  (A, C), then there exists a nonsingular matrix 

Q ∈ R

n×n

such that 

Q−1AQ =⎡⎢⎢⎢⎢⎢⎣

A11   ⋯   A1l

⋮ ⋮Al1   ⋯   All

⎤⎥⎥⎥⎥⎥⎦(11.11a)

with 

Aii =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

0   ⋯   0   ∗1   ⋯   0   ∗⋮ ⋱ ⋮ ⋮0   ⋯   1   ∗

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

ν i×ν i ;   Aij  =

⎡⎢⎢⎢⎢⎢⎣0   ⋯   0   ∗⋮ ⋱ ⋮ ∗0   ⋯   0   ∗

⎤⎥⎥⎥⎥⎥⎦∈ R

ν i×ν j , i ≠  j; (11.11b)

Page 97: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 97/119

EE 580 — Linear Control Systems    8

and 

CQ =

⎡⎢⎢⎢⎢⎢⎣C11   ⋯   C1l

⋮ ⋮Cl1   ⋯   Cll

⎤⎥⎥⎥⎥⎥⎦(11.11c)

with 

Cii = 0   ⋯   0 1 ∈ R1×ν i; (11.11d)

Cij  = 0   ⋯   0 0 ∈ R1×ν i , i <  j;   Cij  = 0   ⋯   0   ∗ ∈ R

1×ν i, i > j.   (11.11e)

(The matrix   Q−1B   does not have any particular structure.)

Proof . The proof is similar to that of Theorem 11.3: Define  σk = ∑ki=1 ν i,  k  = 1, . . . ,  l, and let  pk

be the  σk-th column of  N−1, where  N   is as in (11.10); then use  p1, . . . ,  pl  to form

Q = p1   Ap1   ⋯   Aν 1−1p1   ⋯   pl   Apl   ⋯   Aν l−1pl .   ◻

Suppose that  Q  is as in Theorem 11.7; suppose rankO =  n  and rank C = l. If  ν 1, . . . ,  ν l  are theobservability indices of  (A, C)  with ∑l

i=1 ν i =  n, let

Ao = ⎡⎢⎢⎢⎢⎢⎣A1   0

⋱0   Al

⎤⎥⎥⎥⎥⎥⎦∈ R

n×n with   Ai = ⎡⎢⎢⎢⎢⎢⎢⎢⎣

0   ⋯   0 01   ⋯   0 0⋮ ⋱ ⋮ ⋮0   ⋯   1 0

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ R

ν i×ν i, i = 1, . . . , l; (11.12a)

Co =

⎡⎢⎢⎢⎢⎢⎣C1   0

⋱0   Cl

⎤⎥⎥⎥⎥⎥⎦∈ R

l×n with   Ci = 0   ⋯   0 1 ∈ R1×ν i , i = 1, . . . , l.   (11.12b)

Then we haveQ−1AQ =  Ao + A p

 Co   and   CQ = C p Co   (11.12c)

for some  A p ∈ Rn×l and  C p ∈ R

l×l. If  p1, . . . ,  pl  and  Q  are as in the proof of Theorem 11.7, then itcan be shown that

A p =  Q−1 Aν 1p1   ⋯   Aν lpl   and   C p =  C Aν 1−1p1   ⋯   Aν l−1pl .

Furthermore, the matrix   C p   is in fact a lower triangular matrix with ones on the diagonal, andhence it is nonsingular. The following is the observable version of the   structure theorem .

Theorem 11.8   If  Q  is as in Theorem 11.7, and if  A p  and  C p  are as in   (11.12), then the transfer  function matrix of the system   (11.3)  is given by   H(s) = D(s)−1N(s) + D, where 

D(s) = Λ(s) − S(s)A pC−1 p   ,   N(s) = S(s)Q−1B,

Λ(s) =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

sν 1

sν 2

sν l

⎤⎥⎥⎥⎥⎥⎥⎥⎦

,   S(s) = diag 1   s   ⋯   sν 1−1 , . . . , 1   s   ⋯   sν l−1 .

Proof . The proof is similar to that of Theorem 11.4, and the result follows from D(s)CQ =S(s)sI − Q−1AQ.   ◻In the example presented at the end of Section 11.2.2, a state-space model (A, B, C, D)   was

given, and its controller-form representation (Ac, Bc, Cc, D) and transfer function matrix  H(s) wereobtained. Now, suppose (  A,  B,  C,  D) = (AT, CT, BT, DT). Then an observer-form representationof  (  A,  B,  C,  D) turns out to be (AT

c , CTc , BT

c , DT). Moreover, by the structure theorem, the transferfunction matrix of  (  A,  B,  C,  D)   is equal to  H(s)T.

Page 98: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 98/119

EE 580 — Linear Control Systems    9

11.4 Relation Between Stability, Controllability, and Observability

Controllability (or, more precisely, reachability) and observability are dual notions. Each resultconcerning observability can be obtained from its controllability counterpart, and vice versa, byinvoking the following  duality   result.

Theorem 11.9   Let  A ∈ Rn×n,  B ∈ Rn×m, and  C ∈ Rl×n. The following hold:

(a) The matrix   A  is (asymptotically) stable if and only if  AT is (asymptotically) stable;

(b) The pair  (A, B)  is controllable if and only if  (AT, BT)   is observable;

(c) The pair  (A, C)   is observable if and only if  (AT, CT)  is controllable.

Proof . Part (a) follows from the fact that   A   and   AT share a common set of eigenvalues.Let  C  and  O be the controllability and observability matrices, respectively, of the triple (A, B, C).Part (b) follows from the fact that   C   has full row rank if and only if   CT has full column rank.Similarly, part (c) holds because  O  has full column rank if and only if  OT has full row rank.   ◻

We know that asymptotically stable systems are BIBO stable, but that the converse does nothold true in general. In view of Kalman’s decomposition theorem, it seems that the converseshould hold true if the system is both controllable and observable. This is indeed the case due tothe following result proved in [1, Theorem 12.8].

Theorem 11.10   Suppose the system   (11.3)   is controllable and observable. Then the system is bounded-input bounded-output stable if and only if it is asymptotically stable.

The following result facilitates further insights on the relation between stability, controllability,and observability. These structural properties are connected via Lyapunov equations.

Theorem 11.11   Let  A ∈ Rn×n,  B ∈ R

n×m, and  C ∈ Rl×n. For continuous-time LTI systems of the 

 form   (11.3)  the following hold:(a) Suppose  (A, B)   is controllable. Then   A   is asymptotically stable (i.e., all eigenvalues of   A

have negative real parts) if and only if there exists  P ∈ Rn×n such that 

P > 0   and    AP + PAT = −BBT.

(b) Suppose  (A, C)   is observable. Then   A   is asymptotically stable if and only if there exists Q ∈ R

n×n such that Q > 0   and    ATQ + QA = −CTC.

For discrete-time LTI systems of the form 

x(t + 1) =  Ax(t) + Bu(t),

y(t) =  Cx(t) + Du(t),

the following hold:

(c) Suppose  (A, B)   is reachable. Then  A  is asymptotically stable (i.e., all eigenvalues of  A have magnitudes less than one) if and only if there exists  P ∈ R

n×n such that 

P > 0   and    APAT − P = −BBT.

Page 99: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 99/119

EE 580 — Linear Control Systems    10

(d) Suppose  (A, C)   is observable. Then   A   is asymptotically stable if and only if there exists Q ∈ R

n×n such that Q > 0   and    ATQA − Q = −CTC.

Proof . To prove part (a), suppose AP + PAT = −BBT for some P  >  0. Choose any eigenvalue  λ

of  A, and let  v  be the corresponding eigenvector of  AT. Then

 (λ∗ + λ

)v∗Pv =  v∗

(AP + PAT

)v =

−v∗BBTv. Since  P  >  0  (and  v  ≠ 0), we have  v∗Pv > 0. Also, since (A, B)   is controllable, we haveBTv ≠ 0 by the PBH eigenvector test for controllability, so  v∗BBTv > 0. Thus, R(λ) = (λ∗+ λ)2 =−v∗BBTv(2v∗Pv) < 0. Conversely, supp ose every eigenvalue of  A has negative real part. Let P(t)be the observability Gramian of  (AT, BT) on [0, t]; that is, P(t) = ∫  t0  eAτ BBTeA

Tτ dτ . Since (A, B)is controllable, (AT, BT) is observable, so P(t) >  0  for all  t  > 0 and  P(t) ≥  P(s) whenever t  ≥  s > 0.On the other hand, since   A   is exponentially stable in continuous time, there are   c,   λ  >  0 suchthat eAτ  ≤  ce−λτ , so we have P(t) ≤  c2BBT(2λ) < ∞. Therefore  P = ∫  ∞0   eAτ BBTeA

Tτ dτ   iswell-defined and  P >  0. Moreover,

AP + PAT =    ∞0

  d

dτ eAτ BBTeA

Tτ  + eAτ BBT   d

dτ  eA

Tτ dτ 

=    ∞0

ddτ  eAτ BBTeA

Tτ dτ  = −BBT.

This establishes part (a). Part (b) is a dual result, which follows from part (a) applied to (AT, CT).To prove part (c), suppose   APAT − P  =   −BBT for some   P   >   0. Choose any eigenvalue   λ

of  A, and let  v  be the corresponding eigenvector of  AT. Then proceeding as in part (a) we obtainλ∗λ − 1 =  −v∗BBTvv∗Pv <  0, which implies   λ < 1. Conversely, suppose the magnitude of everyeigenvalue of  A is less than one. Let  P(K ) = ∑K −1

τ =0   Aτ BBT(AT)τ ; that is, let  P(K ) be the discrete-time observability Gramian of  (AT, BT)  on  {0, 1, . . . , K  }. Since  (A, B)  is reachable, we have that(AT, BT)  is observable, and that  P(K ) > 0  and  P(K  + 1) ≥  P(K )  for all positive integers  K . Onthe other hand, since  A  is exponentially stable in discrete time, there are  c > 0 and  λ ∈ (0, 1)  suchthat   Aτ  ≤  cλτ , and so   P(K ) ≤  c2BBT(1 − λ2) < ∞. Therefore  P =

 ∑∞τ =0 Aτ BBT(AT)τ  is

well-defined and  P  > 0. Moreover,  APAT − P = (P − BBT) − P = −BBT. This establishes part (c).Part (d) is dual to part (c), and follows from part (c) applied to  (AT, CT).   ◻

References

[1] W. J. Rugh, Linear System Theory . Englewood Cliffs, NJ: Prentice-Hall, 1993.

[2] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[3] ——,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

Page 100: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 100/119

EE 580 — Linear Control Systems

XII. Eigenvalue Assignment and Stabilization 

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

12.1 Introduction

Recall that a control system is a feedback interconnection of two dynamical systems, called theplant and the controller, as shown in Fig. 12.1. By closing the feedback loop, one aims to make the

controlled output  z exhibit desired properties with respect to the reference input  r.The single most important goal of feedback, or  closed-loop, control is to cope with uncertainties.

One of the very first control problems we have encountered is the controllability problem, wherethe control objective is to steer the initial state to a desired state. A particular solution u  to thisproblem has been given as a function of time that is determined solely by the initial state as wellas known system coefficients. However, this kind of control schemes, so-called open-loop   control,is not satisfactory in real-world situations that involve uncertainties in the initial state, systemcoefficients, etc. What is required to make the control system perform satisfactorily in practiceis to update the control input   u(t) based on the information (about the plant) available (to thecontroller) at each time  t  using a   feedback control law   γ  of the form  u(t) = γ (y(s), t0 ≤  s  ≤  t).

Restricting our attention to linear time-invariant feedback control laws, we describe  γ  by

xK (t) = AK xK (t) + BK y(t) + Fr(t),

u(t) = CK xK (t) + DK y(t) + Gr(t).

Controllers of this form are called  dynamic output feedback controllers . A dynamic output feedbackcontroller has to address two problems simultaneously: One is the problem of transferring the stateof the plant to a desired value (i.e., the control problem), and the other is the problem of estimatingthe state of the plant based on the available information (i.e., the estimation problem). For manycontrol objectives (e.g., eigenvalue assignment and stabilization), however, these two problems areseparated. That is, to design a dynamic output feedback controller, one may design a state feedbackcontroller and a state observer separately and combine them in a straightforward manner.

z Plantr

u+

+

Controller 

yPre-Filter 

Figure 12.1: A feedback control system

1

Page 101: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 101/119

EE 580 — Linear Control Systems    2

12.1.1 State Feedback Controller

The problem of state feedback control is to find a linear static control law of the form

u(t) = Kx(t)

for the plant  x(t) = Ax(t) + Bu(t) with  y(t) = x(t). Closing the feedback loop then gives

x(t) = (A + BK)x(t).

For example, the state feedback stabilization problem is to find a linear  state feedback gain matrix  Ksuch that the closed-loop plant state matrix  A + BK  is asymptotically stable.

12.1.2 State Observer

A state estimation problem is to obtain an estimate  x of the state  x  of the plant using a dynamicalsystem, called the  Luenberger observer , of the form

˙x(t) = Ax(t) + Bu(t) + L

Cx(t) + Du(t) − y(t)

.

If we let the estimation error at time   t   be   e(t) =   x(t) −  x(t), and if the plant is described byx(t) =  Ax(t) +  Bu(t) and  y(t) =  Cx(t) +  Du(t), then the problem of designing a state observerreduces to that of obtaining the matrix  L  such that

e(t) = (A + LC)e(t).

For example, an asymptotic state estimation problem is to find a linear   output injection gain matrix   L  such that the observer state matrix  A + LC  is asymptotically stable.

12.1.3 Dynamic Output Feedback Controller

As we will see shortly, the separation principle for eigenvalue assignment states that one can assign

the eigenvalues of  A + BK and  A + LC separately to place the closed-loop system poles at desiredlocations in the complex plane C. Similarly, the separation principle for stabilization states that, toasymptotically stabilize the closed-loop system, one only needs to obtain a stabilizing state feedbackgain and a stabilizing output injection gain separately. Moreover, if  K  and  L  are the state feedbackand output injection gains obtained separately, then the coefficients   AK ,   BK ,   CK , and   DK   of adynamic output feedback controller can be given in terms of  K  and  L. Lastly,  F  and  G  are chosento satisfy additional performance requirements.

12.2 Eigenvalue Assignment

Consider the linear time-invariant system

x(t) = Ax(t) + Bu(t),y(t) = Cx(t) + Du(t),

  (12.1)

where   A  ∈  Rn×n,   B  ∈  R

n×m,   C  ∈  Rl×n, and   D  ∈   R

l×m. We first study the problem of state-feedback eigenvalue assignment.

Lemma 12.1   If   (A, B)   is controllable and if   rank B  =  m, then by state feedback   u(t) =   Kx(t)with   K   ∈   Rm×n, the eigenvalues of   A +  BK   can be arbitrarily assigned provided that complex eigenvalues appear in conjugate pairs.

Page 102: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 102/119

EE 580 — Linear Control Systems    3

Proof . Let  Q ∈ Rn×n be such that (Ac, Bc) = (Q−1AQ, Q−1B) is in controller form; that is,

Ac  =

A11   · · ·   A1m...

  ...Am1   · · ·   Amm

  and   Bc  =

B11   · · ·   B1m...

  ...Bm1   · · ·   Bmm

with

Aii =

0 1   · · ·   0...

  ...  . . .

  ...0 0   · · ·   1∗ ∗ · · · ∗

 ∈ Rµi×µi ;   Aij  =

0   · · ·   0...

  . . .  ...

0   · · ·   0∗ · · · ∗

 ∈ Rµi×µj , i = j ;

Bii  =

0...01

 ∈ Rµi×1;   Bij  =

0...0∗

 ∈ Rµi×1, i < j;   Bij  =

0...00

 ∈ Rµi×1, i > j,

where  µ1, . . . ,  µm  are the controllability indices of (A, B). Let  λ1, . . . ,  λn  be desired eigenvalueswith complex conjugate eigenvalues appearing in pair, so that the desired characteristic polynomial

 p(s) =n

i=1

(s − λi) = sn + β n−1sn−1 + · · · + β 1s + β 0

for some   β 0, . . . ,   β n−1   ∈   R. Since rankB   =   m, we have   µi   ≥   1 for all   i   = 1, . . . ,   m, so   Bc

has exactly  m  nonzero rows. These rows form an  m-by-m   nonsingular matrix  Bq, which is uppertriangular with ones on the diagonal. Because the corresponding rows of  Ac  denoted by strings of  ∗form an  m-by-n matrix  Aq, the corresponding rows of  Ac + BcKc, forming an  m-by-n matrix  Ad,can be arbitrarily assigned by a suitable choice of  Kc  ∈  Rm×n. That is, for any  Ad  ∈  Rm×n, theunique solution to   Aq  + BqKc   =   Ad   is given by   Kc   =   B−1

q   (Ad − Aq). In particular, we maychoose  Kc  such that

Ac + BcKc  =

0 1   · · ·   0...

  ...  . . .

  ...0 0   · · ·   1

−β 0   −β 1   · · · −β n−1

.

The characteristic polynomial of   Ac  +  BcKc   equals   p(s). Then, it follows from   Ac  +  BcKc   =Q−1

A +  BKcQ−1

Q   that, with   K  =   KcQ−1, the matrix   A +  BK  has the same characteristic

polynomial and hence the desired eigenvalues.  

Theorem 12.2  By state feedback  u(t) = Kx(t) with  K  ∈Rm×n

, the eigenvalues of  A + BK can be arbitrarily assigned if and only if  (A, B)   is controllable, provided that complex eigenvalues appear in conjugate pairs.

Proof . To show sufficiency, suppose (A, B) is controllable. If rank B   =   m, then arbitraryeigenvalue assignment is possible by Lemma 12.1. If rank B   =  r < m, on the other hand, thenthere exist a full-rank matrix  Br  ∈ Rn×r and a nonsingular matrix  F ∈ Rm×m such that

BF =

Br   0

.   (12.2)

Page 103: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 103/119

EE 580 — Linear Control Systems    4

The pair (A, Br) is controllable because the columns of the controllability matrix of (A, B) arelinear combinations of the columns of the controllability matrix of (A, Br). Thus, by Lemma 12.1,there exists a  Kr  ∈ R

r×n such that  A + BrKr  has desired eigenvalues. However, if we let

K =  F

Kr

0

,

then we have

A + BK =  A + BF

Kr

0

 =  A +

Br   0

Kr

0

 =  A + BrKr,

so  A + BK has the desired eigenvalues. This proves the sufficiency part of desired result.To show necessity, suppose (A, B) is not controllable. If  nr denotes the rank of the controllability

matrix of (A, B), then we have  nr  < n. Thus there exists a nonsingular  Q ∈  Rn×n such that, nomatter what state feedback gain matrix  K ∈ Rm×n is used, we have

Q−1(A + BK)Q =  Q−1AQ + Q−1B(KQ) =

A1   A12

0 A2

+

B1

0

K1  K2

= A1 + B1 K1   A12 + B1 K2

0 A2 ,

where A1 ∈ Rnr×nr and B1 ∈ R

nr×m are such that (A1, B1) is controllable, and where K1 ∈ Rm×nr

and K2 ∈ Rm×(n−nr) are such that

K1  K2

 =  KQ. The eigenvalues of  A + BK are the same as

the eigenvalues of  Q−1(A + BK)Q, which in turn are the eigenvalues of  A1 + B1K1  and those of 

A2   combined. Since the eigenvalues of   A2  are unaffected by state feedback, arbitrary eigenvalueassignment is not possible. This shows necessity, and completes the proof.  

According to Theorem 12.2, an exact condition for arbitrary eigenvalue assignment with statefeedback is the controllability of the pair (A, B). In general, the proof of the theorem suggeststhat all controllable eigenvalues are arbitrarily assignable via state feedback and none of the un-controllable eigenvalues is affected by state feedback. The problem of output-injection eigenvalue

assignment can be addressed by invoking duality.

Theorem 12.3  By output injection with   L ∈  Rn×l, the eigenvalues of   A + LC  can be arbitrarily assigned if and only if  (A, C)   is observable, provided that complex eigenvalues appear in conjugate pairs.

Proof . The result is immediate from Theorem 12.2, the duality result, and the fact that A + LC

and

A + LCT

share the same set of eigenvalues.  

12.3 Example of Eigenvalue Assignment

Let

A = 0 1 0

0 0 10 2   −1

  and   B = 0 1 1

1 1 00 0 0

.

Since rank B = 2 <  3, there are a full-rank matrix  Br  and a nonsingular matrix  F  such that (12.2)holds. In particular, because the three columns b1,  b2,  b3  of  B  are such that  {b1, b2}  is linearlyindependent and such that  b1 − b2 + b3 = 0, we may take

Br  =

0 11 10 0

  and   F =

1 0 10 1   −10 0 1

.

Page 104: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 104/119

EE 580 — Linear Control Systems    5

The controllability matrix of the pair (A, Br) is

Cr  =

0 1 1 1 0 01 1 0 0 2 20 0 2 2   −2   −2

with rank

Cr  = 3, so (A, Br) as well as (A, B) is controllable. The first three columns of 

  Cr   arelinearly independent, so

µ1 = 2, µ2 = 1,   M =

b1   Ab1   b2

 =

0 1 11 0 10 2 0

,   and   M−1 =

−1 1 1/20 0 1/21 0   −1/2

.

From  σ1  = µ2  = 2 and  σ2  = µ1 + µ2  = 3, it follows that  q1  and  q2  are the second and third rowsof  M−1, respectively:

q1 =

0 0 1/2

  and   q2 =

1 0   −1/2

.

Then the similarity transformation  Q determined by

Q−

1 = q1

q1Aq2

 = 0 0 1/2

0 1   −1/21 0   −1/2

  ⇒   Q = 1 0 1

1 1 02 0 0

gives us the controller canonical form (Ac, Bc) of (A, Br), where

Ac  =  Q−1AQ =

0 1 02   −1 01 0 0

  and   Bc  =  Q−1Br =

0 01 10 1

.

Suppose that we want to assign the closed-loop eigenvalues to the locations  −2, −1 + i, and −1 − i;that is, suppose that the desired eigenvalues are the roots of the polynomial

s3 + β 2s2 + β 1s + β 0 = (s + 2)(s2 + 2s + 2) = s3 + 4s2 + 6s + 4.

Taking the second and third rows of  Ac  and  Bc  yields the following desired equality:2   −1 01 0 0

+

1 10 1

Kc  =

  0 0 1

−β 0   −β 1   −β 2

 =

 0 0 1−4   −6   −4

.

This linear equation has a unique solution given by

Kc  =

1 10 1

−1

 0 0 1−4   −6   −4

2   −1 01 0 0

 =

 3 7 5−5   −6   −4

.

(By the way, this does not mean that  Kc  is the only state feedback gain that achieves the desiredeigenvalue assignment—see [1, Example 9.11].) Because   Ac  +  BcKc  has the desired eigenvalues,A + BrKr  has the desired eigenvalues as well, where

Kr  =  KcQ−1 =  5 7   −9/2

−4   −6 5/2

.

Finally, the state feedback gain

K =  F

Kr

0

 =

5 7   −9/2−4   −6 5/20 0 0

  ⇒   A + BK =

−4   −5 5/21 1   −10 2   −1

,

where the eigenvalues of  A + BK  are indeed  −2, −1 + i, and −1 − i.

Page 105: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 105/119

EE 580 — Linear Control Systems    6

12.4 Stabilization

12.4.1 Stabilizability and Detectability

Consider the linear time-invariant system (12.1) with   A   ∈   Rn×n,   B   ∈   Rn×m,   C   ∈   Rl×n, andD ∈ Rl×m.

Definition 12.4   The pair   (A, B)   is said to be   stabilizable   if there exists a state feedback gain matrix  K ∈ Rm×n such that  A + BK   is asymptotically stable.

Theorem 12.5  The pair  (A, B) is stabilizable if and only if al l the uncontrollable eigenvalues of  Ahave negative real parts.

Proof . Suppose that all the uncontrollable eigenvalues of  A  have negative real parts. Let  Q  bea nonsingular matrix such that (A, B) is in the standard form for uncontrollable systems:

A =  Q−1AQ =

A1   A12

0 A2

  and   B =  Q−1B =

B1

0

,   (12.3)

where (A1, B1), with   A1   ∈   Rnr×nr and   B1   ∈   Rnr×m, is controllable but all the eigenvalues of A2   ∈   R(n−nr)×(n−nr) are uncontrollable and hence have negative real parts. By Theorem 12.2,there exists a   K1  ∈  Rm×nr such that all the eigenvalues of   A1 +  B1K1  have negative real parts.Let

K =

K1   0

 ∈ Rm×n.

Then all the eigenvalues of the matrix

A + BK =

A1   A12

0 A2

+

B1

0

K1   0

 =

A1 + B1K1   A12

0 A2

,

which is of block triangular form, have negative real parts. The matrices   A +  BK   and   Q(A +

BK)Q−1 share the same set of eigenvalues, but  Q(A + BK)Q−1 = A + BKQ−1. Thus

K =  KQ−1 =

K1   0

Q−1

is a stabilizing state feedback gain matrix such that  A + BK  is asymptotically stable.Conversely, suppose that (A, B) is stabilizable, so that  A+BK is asymptotically stable for some

K ∈  Rm×n. Let  Q  represent any similarity transformation such that (A, B) is in the form (12.3),where the pair (A1, B1), with A1 ∈ Rnr×nr and B1 ∈ Rnr×m, is controllable and all the eigenvaluesof  A2  ∈  R

(n−nr)×(n−nr) are uncontrollable. Let [ K1 K2], with K1  ∈  Rm×nr and K2  ∈  Rm×(n−nr),

be a partition of  KQ. Then, since the set of eigenvalues of 

Q

−1

(A + BK)Q =  A + BKQ = A1 + B1 K1   A12 + B1 K2

0 A2 is equal to that of  A+BK, the eigenvalues of  A2 have negative real parts. That is, all uncontrollableeigenvalues have negative real parts. This completes the proof.  

The above theorem says that there exists a stabilizing state feedback gain for a linear time-invariant system if and only if uncontrollable eigenvalues of the system have negative real parts.An immediate consequence is that controllable linear time-invariant systems are stabilizable.

Corollary 12.6   The pair  (A, B)  is stabilizable if it is controllable.

Page 106: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 106/119

EE 580 — Linear Control Systems    7

Due to duality, all these results on stabilizing state feedback can be restated in terms of stabi-lizing output injection.

Definition 12.7   The pair   (A, C)   is said to be   detectable   if there exists an output injection gain matrix  L ∈ R

n×l such that  A + LC  is asymptotically stable.

Theorem 12.8   The pair   (A, C)   is detectable if and only if all the unobservable eigenvalues of  Ahave negative real parts.

Corollary 12.9   The pair  (A, C)  is detectable if it is observable.

12.4.2 Separation Principle

For simplicity, suppose a linear time-invariant plant (12.1) has   D  =   0. Also, since the referenceinput is irrelevant to stabilization, assume  r   is identically zero. We will consider dynamic outputfeedback controllers of the form

xK (t) = AK xK (t) + BK y(t),

u(t) = CK xK (t) + DK y(t),

  (12.4)

where  AK   ∈  RnK×nK ,   BK   ∈  R

nK×l,   CK   ∈  Rm×nK , and  DK   ∈  R

m×l. Closing the feedback loopyields the homogeneous closed-loop system

 x(t)xK (t)

 =

A + BDK C BCK 

BK C AK 

 x(t)xK (t)

.   (12.5)

Theorem 12.10  There exists a dynamic output feedback controller   (12.4) such that the closed-loopsystem  (12.5) is asymptotically stable if and only if the triple  (A, B, C) is stabilizable and detectable.Moreover, if  K  ∈ Rm×n and  L  ∈ Rn×l are such that  A + BK and  A + LC are asymptotically stable,then a stabilizing dynamic output feedback controller can be taken to be with  nK  = n  and 

AK  = A + LC + BK,   BK  = −L,   CK  = K,   and    DK  = 0.   (12.6)

Proof . Suppose that the closed-loop system (12.5) is asymptotically stable. Let

A =

A 0

0 0

 ∈ R

(n+nK)×(n+nK),   B =

0 B

I 0

 ∈ R

(n+nK)×(nK+m),

C =

0 I

C 0

 ∈ R

(nK+l)×(n+nK),   and  K =

AK    BK 

CK    DK 

 ∈ R

(nK+m)×(nK+l).

Then (12.5) can be written as˙x(t) =

A +

 B

K

C

x(t),

where x(t) = [x(t)T xK (t)T]T ∈ Rn+nK is the closed-loop state. Thus A + BKC being asymptoti-cally stable implies that A, B, C  is both stabilizable and detectable. We have

rank

λI − A B = rank

λI − A 0 0 B

0   λI I 0

 = rank

λI − A B

+ nK ;

rank

λI − AC

 = rank

λI − A 0

0   λI

0 I

C 0

 = rank

λI − A

C

+ nK 

Page 107: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 107/119

EE 580 — Linear Control Systems    8

for all λ ∈ C. Moreover, every eigenvalue of  A  is an eigenvalue of  A. Thus, by the PBH rank test,we conclude that (A, B, C) is both stabilizable and detectable as well. This proves the necessitypart of the desired result.

Conversely, suppose that (A, B, C) is both stabilizable and detectable, so that there exist   K

and  L  such that  A + BK  and  A + LC are asymptotically stable. If the dynamic output feedback

controller (12.4) have  n  =  nK  and (12.6), then we haveI 0

I   −I

−1

A + BDK C BCK 

BK C AK 

I 0

I   −I

 =

A + BK   −BK

0 A + LC

,

and hence the closed-loop system (12.5) is asymptotically stable. This shows the sufficiency partholds true as well.  

Corollary 12.11  There exists a dynamic output feedback controller  (12.4) such that the closed-loopsystem   (12.5)  is asymptotically stable if the triple  (A, B, C)   is controllable and observable.

12.5 Observer-Based Controllers

12.5.1 Full-Order Observers

It is readily seen that the controller structure in (12.6) corresponds to an  observer-based controller of the form

˙x(t) = Ax(t) + Bu(t) + L

Cx(t) + Du(t) − y(t)

,   (12.7a)

u(t) = Kx(t) + Gr(t) (12.7b)

applied to the plant (12.1) (with possibly nonzero  D). The separation principle in Theorem 12.10says that, as far as dynamic output feedback stabilization is concerned, using an observer-basedcontroller suffices. The order of the observer (12.7a) (i.e., the number of the observer state variablesin  x) is equal to  n, which is the same as the order of the plant. For this reason, this observer iscalled a  full-order observer . The state estimation error  e =  x − x  satisfies

e(t) = (A + LC)e(t),

so there exists an asymptotically stable state observer of the form (12.7a) for the plant (12.1) if and only if the pair (A, C) is detectable. With the closed-loop state defined by [xT eT]T, thestate-space description of the closed-loop system reads

x(t)e(t)

 =

A + BK   −BK

0 A + LC

x(t)e(t)

+

BG

0

r(t),

y(t) =

C + DK   −DK x(t)

e(t)

+ DGr(t),

so the closed-loop transfer function matrix from  r  to  y  is given by

 H(s) = C + DK   −DK (sI − A − BK)−

1 ∗0   (sI − A − LC)−1

BG0 + DG

= (C + DK)(sI − A − BK)−1BG + DG.

The observer eigenvalues uncontrollable, so they do not manifest themselves as system poles.This suggests how one can address certain model matching problems via observer-based control:If (A, B, C) is controllable and observable, then choose  K  to place the system poles at desired lo-cations and choose  L  such that the observer eigenvalues are asymptotically stable and much fasterthan the system poles so as not to affect the transient behavior of the overall system much.

Page 108: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 108/119

EE 580 — Linear Control Systems    9

12.5.2 Reduced-Order Observers

Suppose that the output matrix   C   has full row rank (i.e., rank C   =   l). Let   Q   ∈   Rn×n be anonsingular matrix whose inverse is of the form

Q−1 = C

R .   (12.8)

Then we haveCQ =

Il   0

,

where  Il   is the  l-by-l   identity matrix. Then the similarity transformation  Qx(t) = x(t) gives

˙x(t) = Q−1AQx(t) + Q−1Bu(t) =

A11   A12

A21   A22

x1(t)x2(t)

+

B1

B2

u(t),

y(t) = CQx(t) + Du(t) =

I 0 x1(t)

x2(t)

+ Du(t).

At each time   t, we have   y(t) −  Du(t) =  x1(t) with   u(t) and   y(t) known, so the vector  x1(t) is

perfectly observed and only  x2(t) needs to be estimated. If   D  =   0, assume   u(t) is continuouslydifferentiable in  t. Putting

u(t) = A21x1(t) + B2u(t) = A21y(t) + (B2 − A21D)u(t),   (12.9a)

y(t) =   ˙x1(t) − A11x1(t) − B1u(t) =  y(t) − Du(t) − A11y(t) − (B1 − A11D)u(t),   (12.9b)

we obtain

˙x2(t) = A22x2(t) + u(t),

y(t) = A12x2(t).  (12.10)

Therefore, the problem of estimating the  n-dimensional state vector  x(t) of the system (A, B, C),where   C  has full row rank, has been reduced to that of estimating the ( n − l)-dimensional state

vector  x2(t) of the system (A22, I, A12).Observability is invariant under the order-reduction described above. This invariance property

is a consequence of the PBH test for observability:

Lemma 12.12   The pair  (A22, A12)   is observable if and only if  (A, C)  is observable.

Proof . If (A, C) is observable, then so is (Q−1AQ, CQ). This implies thatλI − A11   −A12

−A21   λI − A22

I 0

v1

v2

 = 0 (12.11)

holds for some λ  ∈ C only if  v1  = 0 and v2 = 0. In particular, if  v1 = 0 and if  λ  and  v2 = 0 are such

that  A22v2  = λv2, then we must have  A12v2  = 0. Hence (A22, A12) is observable. Conversely, if (A, C) is not observable, then there exist  λ ∈  C,  v1  ∈  Cl and  v2  ∈  C(n−l) with

vT1   vT

2

 = 0 such

that (12.11) holds. Since (12.11) implies  v1  = 0, we must have that   v2   is an eigenvector of   A22

such that  A12v2  = 0. Hence (A22, A12) is not observable.  

Suppose that (A, C) is observable, and that   L  ∈  R(n−l)×l is an output injection gain matrix

such that the eigenvalues of the closed-loop matrix  A22 + LA12  have negative real parts. Then anasymptotically stable state observer for the system (12.10) is of the form

˙x2(t) = A22x2(t) + u(t) + L(A12x2(t) − y(t)),   (12.12)

Page 109: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 109/119

EE 580 — Linear Control Systems    10

where x2(t) is the estimate of  x2(t). Plugging (12.9) into (12.12) and letting

z(t) = x2(t) + L(y(t) − Du(t))

gives

z(t) =(A22 + LA12)z(t) + (B2 + LB1)u(t)+

(A21 + LA11) − (A22 + LA12)L

(y(t) − Du(t)),  (12.13a)

which is a dynamical system of order  n−l that generates an estimate  z(t)−L(y(t)−Du(t)) of  x2(t).Now, from  y(t) = x1(t) + Du(t), it follows that

  y(t) − Du(t)z(t) − L(y(t) − Du(t))

is an estimate of  x(t), and hence that

x(t) = C

R−1

  y(t) − Du(t)

z(t) − L(y(t) − Du(t))   (12.13b)

is an estimate of  x(t). In summary, we have the following result:

Theorem 12.13   Suppose that   (A, C)   is detectable, and that   C  has full row rank (i.e.,   rank C =l). Let   R   ∈   R

(n−l)×n be any matrix such that the right-hand side of    (12.8)   is nonsingular; let L  ∈   R

(n−l)×l be such that all the eigenvalues of   A22  +  LA12   have negative real parts. Then an asymptotically stable state observer of order  n − l   for the plant   (12.1)  is given by   (12.13).

12.6 Example of Observer-Based Control

Let

A = 0 1 0

0 0 10 2   −1

,   B = 0 1 1

1 1 00 0 0

,   C = 2 1 0 ,   and   D = 0 0 0 .

The eigenvalues of   A  are 0, 1, and  −2. The PBH rank test tells us that the eigenvalues 0 and 1are observable and that the eigenvalue  −2 is unobservable:

rank

−A

C

 = rank

I − A

C

 = 3 and rank

−2I − A

C

 = 2.

Since the unobservable eigenvalue has negative real part, the pair (A, C) is detectable and thereexists an asymptotically stable state observer. The unobservable subspace is identified as follows:

O = C

CA

CA2 = 2 1 0

0 2 10 2 1

  ⇒ N (O) = span{v3}   where   v3 = 1

−24 .

Letting   Q   =

e1   e2   v3

, we transform (A, C) to its standard form for unobservable systems

(Ao, Co) = (Q−1AQ, CQ), where

Q =

1 0 10 1   −20 0 4

  ⇒   Ao =

A1   0

A21   A2

 =

0 1/2 00 1 0

0 1/2   −2

,   Co =

C1   0

 =

  2 1 0

.

Page 110: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 110/119

EE 580 — Linear Control Systems    11

Now, we will shift the observable eigenvalues of  A, which are the eigenvalues of  A1, to the left-half 

plane of  C. With L1  =

l1   l2T

, the characteristic polynomial of  A1 + L1C1   is

 p(s) = s2 − (2l1 + l2 + 1)s + (2l1 − l2).

If the desired locations for the observer eigenvalues are  −2, −2, and −5, then, as the unobservable

eigenvalue of   A   is already at   −2, we need to place the roots of   p(s) at   −2 and   −5. Solving p(s) = (s + 2)(s + 5) for  l1  and  l2  gives  l1 = 1/2 and  l2  =  −9. It follows from

A + LC =  QAoQ−1 + LCoQ−1

= Q

A1   0

A12   A2

+ Q−1L

C1   0

Q−1 = Q

A1 + L1C1   0

A21 + L2C1   A2

Q−1,

where Q−1L =LT

1  LT

2

Twith L1 ∈ R

2×1 and L2 ∈ R1×1, that L2  can be arbitrarily chosen, and

that the observer gain  L ∈ R3×1 is determined by

L =  QL1L2

 =1 0 10 1   −2

0 0 4

1/2−90

 =1/2−9

0

with L2   = 0. If the feedback controller gain  K   is as in Section 12.3, then the following observer-based controller places the closed-loop eigenvalues at −5, −2, −2,  −2, and −1 ± i:

˙x(t) = Ax(t) + Bu(t) + L(Cx(t) + Du(t) − y(t))

=

0 1 00 0 10 2   −1

x(t) +

0 1 11 1 00 0 0

u(t) +

1/2−90

2 1 0

x(t) − y(t)

,

u(t) = Kx(t) + Gr(t)

=

−4   −5 5/21 1   −10 2   −1

x(t) + Gr(t).

The closed-loop transfer function matrix with  G =  I  is then

 H(s) = (C + DK)(sI − A − BK)−1BG + DG

=

2 1 0

s   0 00   s   00 0   s

−4   −5 5/21 1   −10 2   −1

−1

0 1 11 1 00 0 0

=   s2 − 5s + 4s3 + 4s2 + 6s + 4

3s2 − 4s + 7s3 + 4s2 + 6s + 4

2s2 + s + 3s3 + 4s2 + 6s + 4

.

References

[1] P. J. Antsaklis and A. N. Michel,  A Linear Systems Primer . Boston, MA: Birkhauser, 2007.

[2] ——,  Linear Systems , 2nd ed. Boston, MA: Birkhauser, 2006.

Page 111: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 111/119

EE 580 — Linear Control Systems

XIII. State-Space Realization of Linear Systems

Department of Electrical Engineering

Pennsylvania State University

Fall 2010

13.1 Introduction

Let A ∈ Rn×n, B ∈ R

n×m, C ∈ Rl×n, and  D ∈ R

l×m. We know that the  external description   (i.e.,the input-output description) of the linear time-invariant system

x(t) = Ax(t) + Bu(t),

y(t) = Cx(t) + Du(t)  (13.1)

is given by the transfer function matrix

 H(s) = C(sI − A)−1B + D,   (13.2)

which is the Laplace transform of the impulse response matrix

H(τ ) =

CeAτ B + Dδ (τ ) if  τ  ≥ 0;

0   if  τ < 0

of the system (13.1). Now, the problem we want to address is to determine an  internal description (i.e., the state-space description) of a given transfer function. That is, we want to implement asystem that exhibits the desired input-output behavior specified by a transfer function matrix.

13.2 Existence of Realizations

13.2.1 Markov Parameters

A rational fraction of polynomials with real coefficients is called a   rational function . The set of rational functions, denoted by  R(s) and equipped with the usual addition and multiplication, is afield. Thus  R(s)n is a vector space over the field of rational functions, and  R(s)m×n is the set of linear transformations that map   R(s)n into   R(s)m. It follows from (13.2) that transfer functions

of linear time-invariant systems belong to  R(s)m×n. Therefore, without loss of generality, we shallrestrict our attention to rational transfer function matrices.

Definition 13.1   The quadruple   (A, B, C, D)   is said to be a   (state-space) realization   of  H(s)  ∈R(s)l×m if   (13.2)  holds.

If the transfer function matrix H(s) ∈ R(s)l×m of a linear time-invariant system has a Laurentseries expansion  H(s) = H0 + H1s−1 + H2s−2 + · · ·   ,

1

Page 112: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 112/119

EE 580 — Linear Control Systems    2

then the matrices  Hk, k  = 0, 1, ..., are called the  Markov parameters  of the system. The Markovparameters can be determined by

H0  = lims→∞

 H(s)

andHk+1 = lim

s→∞

sk+1 H(s) − H0 − H1s−1 − · · · − Hks−kfor all  k  = 0, 1, .... A condition for a (A, B, C, D) to be a realization of some H(s) can be givenin terms of the Markov parameters.

Lemma 13.2   The quadruple  (A, B, C, D)   is a realization of  H(s) ∈ R(s)l×m if and only if 

H0 =  D   and    Hk+1 =  CAkB, k = 0,   1, . . . .

Proof . If  x(t) = Ax(t), then  sx(s) − x(0) = Ax(s), so (sI − A)−1 is the Laplace transform of eAτ , τ  ≥ 0. Since eAτ  has the series expansion

 ∞

k=0 Akτ k/k!, and since the Laplace transform of τ k/k! is  s−(k+1) for each nonnegative integer  k, we have (sI − A)−1 =

 ∞

k=0 Aks−(k+1). Thus

 H(s) = D + C(sI − A)−1B =  D +∞k=0

CAkBs−(k+1) = D + CBs−1 + CABs−2 + · · ·   .  

13.2.2 Block Controller Form Realization

A rational matrix H(s) ∈ R(s)l×m is said to be proper   if lims→∞ H(s) < ∞ entrywise. It is said to

be   strictly proper   if lims→∞ H(s) = 0.

A straightforward way to write down a state-space realization of a proper rational matrix H(s)is to make a multi-variable analog of the single-input controller form (or the single-output observerform) [1]. Let us write

 H(s) = N(s)/d(s),   (13.3a)

where the denominator polynomial 

d(s) = sr + dr−1sr−1 + · · · + d1s + d0   (13.3b)

is the monic least common multiple of the denominators of the entries of  H(s) and

N(s) = Nrd(s) + Nr−1sr−1 + · · · + N1s + N0.   (13.3c)

(A polynomial is called monic if its term of highest degree has a coefficient of one.) Note thatNr = 0   if  H(s) is strictly proper. Let

A = 0 Im   · · ·   0...

  ...

  . ..

  ...0 0   · · ·   Im

−d0Im   −d1Im   · · · −dr−1Im

∈ Rrm×

rm,   B = 0...0

Im

∈ Rrm×

m,   (13.4a)

C =

N0   N1   · · ·   Nr−1

∈ R

l×rm,   D =  Nr  ∈ Rl×m,   (13.4b)

where  Im  denotes the  m-by-m  identity matrix. Then (A, B, C, D) is a realization of  H(s).

Lemma 13.3   Let  H(s) ∈ R(s)l×m be a proper rational matrix satisfying  (13.3). Then  (A, B, C, D)given by   (13.4)  is a realization of  H(s).

Page 113: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 113/119

EE 580 — Linear Control Systems    3

Proof . Let   Γ(s) = [Γ0(s)T · · ·   Γr−1(s)T]T be the matrix formed by the last   m   columns of (sI − A)−1. Then it follows from

(sI − A)Γ =

0   · · ·   0 IT

that sΓ0(s) = Γ1(s), . . . ,  sΓr−2(s) = Γr−1(s), and d0Γ0(s)+· · ·+dr−2Γr−2(s)+(s+dr−1)Γr−1(s) = I.

Equivalently, we havesΓ0(s) = Γ1(s), s2Γ0(s) = Γ2(s), . . . , sr−1Γ0(s) = Γr−1(s),   and   d(s)Γ0(s) = I.

Thus H(s) = C(sI − A)−1B + D  because

 H(s) = Nr + Nr−1sr−1

d(s) + · · · + N1

s

d(s) + N0

1

d(s)

= Nr + Nr−1Γr−1(s) + · · · + N1Γ1(s) + N0Γ0(s)

=

N0   · · ·   Nr−2   Nr−1

∗ · · · ∗   Γ0(s)...   · · ·

  ...  ...

∗ · · · ∗   Γr−2(s)

∗ · · · ∗   Γr−1(s)

0...0

I

+ Nr.  

The quadruple (A, B, C, D) of the form (13.4) shall be said to be in a   block controller form .How to obtain a block observer form is similar.

Theorem 13.4   There exists a realization of  H(s) ∈ R(s)l×m if and only if  H(s)   is proper.

Proof . If (A, B, C, D) is a realization of  H(s), then Lemma 13.2 implies that the first Markovparameter  H0  = D, and hence that lims→∞

 H(s) <  ∞   entrywise. Conversely, if  H(s) is a properrational matrix, then by Lemma 13.3 there exists a block controller form realization of  H(s).  

13.3 Minimal Realizations

If   A   ∈   R

n×n

,   B   ∈   R

n×m

,   C   ∈   R

l×n

, and   D   ∈   R

l×m

, then   n   is called the  order   of the matrixquadruple (A, B, C, D).

Definition 13.5   An  n-th order realization   (A, B, C, D)  of  H(s) ∈ R(s)l×m is said to be   minimal(or   irreducible) if every realization of  H(s)  is at least of order  n.

The notions of controllability and observability play a crucial role in addressing the problem of minimal realization.

Lemma 13.6   Let  (A1, B1, C1, D1) and  (A2, B2, C2, D2) be realizations of  H(s) ∈ R(s)l×m. Define 

Ci(0) = Bi,   Oi(0) = Ci,   Ci(k + 1) =

Bi   AiBi   · · ·   Ak

i Bi

,   (13.5a)

Oi(k + 1) =

Ci

CiAi

...CiA

ki

  (13.5b)

 for  i = 1,  2  and for  k  = 0,  1, . . . . Then we have 

O1(k1)C1(k2) =  O2(k1)C2(k2)

 for all nonnegative integers  k1   and  k2.

Page 114: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 114/119

EE 580 — Linear Control Systems    4

Proof . Due to Lemma 13.2, the Markov parameters of the two realizations are the same: thatis,  D1  =  D2  and  C1Ak

1B1 =  C2Ak2B2   for  k  = 0, 1, . . . . Thus the result follows from

Oi(k1)Ci(k2) =

CiBi   CiAiBi   · · ·   CiAk2−1i   Bi

CiAiBi   CiA2i Bi   · · ·   CiA

k2i   Bi

.

..

  .

..

  .

..CiA

k1−1i   Bi   CiA

k1i   Bi   · · ·   CiA

k1+k2−2i   Bi

.  

Theorem 13.7  A realization   (A, B, C, D)  of  H(s) ∈  R(s)l×m is minimal if and only if it is both controllable and observable.

Proof . If a realization is not controllable and observable, then, using the Kalman decompositionmethod, one may obtain another realization of lower order that is both controllable and observable.Thus a minimal realization is necessarily both controllable and observable.

To show sufficiency, let (A1, B1, C1, D1) and (A2, B2, C2, D2) be realizations of  H(s); supposethat (A1, B1, C1, D1) is of order n1 and both controllable and observable, and that (A2, B2, C2, D2)is of order  n

2. If  Ci(k) and  Oi(k) are as in (13.5), then by Lemma 13.6 we have  O

1(n

1)C

1(n

1) =

O2(n1)C2(n1). Since (A1, B1, C1, D1) is controllable and observable, we have that  O1(n1) has fullcolumn rank and   C1(n1) has full row rank, so that rank C1(n1) = rankO1(n1) =  n1. Choose anymatrix  M ∈ R

ln1×(l−1)n1 such thatO1(n1)   M

  is nonsingular. Then

O1(n1)   M

C1(n1)0

 =  O1(n1)C1(n1) and rank

C1(n1)

0

 = rankC1(n1) = n1.

Thus we must have

rankO1(n1)C1(n1)

 = rank

O2(n1)C2(n1)

 =  n1.

Since   RC2(k)   ⊂ RC2(n2)   for all   k   ≤   n2, and since the Cayley-Hamilton theorem implies

RC2(k)  =  RC2(n2)   for all  k   ≥  n2, we have rankC2(n1)  ≤  n2. Similarly, since  N O2(k)  ⊃ N O2(n2)

 for all k ≤ n2, and since N 

O2(k)

 =  N 

O2(n2)

 due to the Cayley-Hamilton theorem,

we have rankO2(n1)  ≤  n2   by the fundamental theorem of linear equations. Thus we must haverank

O2(n1)C2(n1)

≤ n2. This implies  n1 ≤ n2  and establishes sufficiency.  

Theorem 13.7 suggests that one way to obtain a minimal realization of a proper transfer functionmatrix H(s) is to find any realization of  H(s) (e.g., a block controller form realization) and thenextract the controllable and observable part of it using the Kalman canonical decomposition. Thefollowing result characterizes all minimal realizations.

Theorem 13.8   Let   (A1, B1, C1, D1)   be a given minimal realization of 

 H(s)   ∈   R(s)l×m. Then 

(A2, B2, C2, D2)   is another minimal realization of 

 H(s)   if and only if there exists a nonsingular 

matrix  P ∈ R

n×n

such that 

A2 = P−1A1P,   B2  =  P−1B1,   C2 =  C1P,   and    D2 = D1.   (13.6)

Moreover, such a matrix  P  is given by 

P =  C1CT2

C2C

T2

−1

=O

T1O1

−1O

T1O2,

where   Ci   and  Oi   are the controllability and observability matrices of  (Ai, Bi, Ci, Di), respectively, for  i = 1,  2.

Page 115: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 115/119

EE 580 — Linear Control Systems    5

Proof . If (13.6) holds true for some nonsingular   P, then (A2, B2, C2, D2) is equivalent to(A1, B1, C1, D1), and hence it is a realization of  H(s) as well. Moreover, since  P−1A1P  ∈  R

n×n

whenever  A1  ∈ Rn×n, we have that (A2, B2, C2, D2) is of minimal order whenever (A1, B1, C1, D1)

is of minimal order. This proves the sufficiency part of the desired result.To prove necessity, suppose that (A1, B1, C1, D1) and (A2, B2, C2, D2) are minimal realizations

of  H(s). We will show that they are equivalent. Since the two realizations are minimal, they areof the same order, say  n. Due to Lemma 13.6, we have  O1C1  =  O2C2, which gives

OT1O1C1 =  O

T1O2C2   and   O1C1C

T2   =  O2C2C

T2 .   (13.7)

By Theorem 13.7, the two realizations are b oth controllable and observable, so in particular C1CT2 ,

C2CT2 ,  OT

1O1, and  OT1O2  are nonsingular. (This follows from  Sylvester’s rank inequality : If  M ∈

Rm×n and  N  ∈  R

n×k, then rank M + rank N − n  ≤  rank(MN).) Define nonsingular matrices P,

Q ∈ Rn×n by P =  C1C

T2

C2C

T2

−1

and  Q  =O

T1O1

−1O

T1O2. Then (13.7) yields

P =  C1CT2

C2C

T2

−1

=O

T1O1

−1O

T1O2C2C

T2

C2C

T2

−1

=O

T1O1

−1O

T1O2 =  Q.   (13.8)

Let   Ci(k) and  Oi(k) be as in (13.5). Then  OiAiCi  are submatrices of  Oi(n + 1)Ci(n + 1), andhence Lemma 13.6 implies  O1A1C1 =  O2A2C2, which gives

OT1O1A1C1C

T2   =  O

T1O2A2C2C

T2 .   (13.9)

Now (13.7)–(13.9) yield  A1P  =  PA2,   C1  =  PC2, and  O1P  =  O2, from which the equivalence of the two realizations follows.  

13.4 Discrete-Time Case

Let   A   ∈   Rn×n,   B   ∈   R

n×m,   C   ∈   Rl×n, and   D   ∈   R

l×m. The input-output description of thediscrete-time linear time-invariant system

x(t + 1) = Ax(t) + Bu(t),y(t) = Cx(t) + Du(t)

  (13.10)

is given by the transfer function matrix

 H(z) =∞τ =0

z−τ H(τ ) = C(zI − A)−1B + D,   (13.11)

where  H(τ ) is the impulse response matrix of the system (13.10) given by

H(τ ) =

CAτ −1B   if  τ > 0;

D   if  τ  = 0.

Definition 13.9   The quadruple   (A, B, C, D)   is said to be a   realization   of  H(z)   ∈   R(z)l×m if (13.11)  holds.

Theorem 13.10  There exists a discrete-time realization of  H(z) ∈ R(z)l×m if and only if  H(z)  is proper.

Theorem 13.11   A discrete-time realization   (A, B, C, D)   of  H(z)   ∈   R(z)l×m is minimal if and only if it is both reachable and observable.

Page 116: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 116/119

EE 580 — Linear Control Systems    6

Theorem 13.12   Let  (A1, B1, C1, D1) be a given minimal realization of  H(z) ∈ R(z)l×m in discrete time. Then  (A2, B2, C2, D2)   is another minimal realization of  H(z)   in discrete time if and only if there exists a nonsingular matrix  P ∈ R

n×n such that 

A2 = P−1A1P,   B2  =  P−1B1,   C2 =  C1P,   and    D2 = D1.

13.5 An Example of Minimal Realizations

In this example, we will obtain a minimal realization of 

 H(s) =

1

(s − 1)21

(s − 1)(s + 3)−6

(s − 1)(s + 3)2s − 2

(s + 3)2

by applying the Kalman decomposition theorem on its block controller form realization. The leastcommon multiple of the denominators of the entries of  H(s) is given by

d(s) = sr + dr−1sr−1 + · · · + d1s + d0 = (s − 1)2(s + 3)2

= s4 + 4s3 − 2s2 − 12s + 9,

and we can write H(s) = N(s)/d(s), where

N(s) =

  (s + 3)2 (s − 1)(s + 3)−6(s − 1) (s − 1)2(s − 2)

 =

s2 + 6s + 9   s2 + 2s − 3

−6s + 6   s3 − 4s2 + 5s − 2

= N4d(s) + N3s3 + N2s2 + N1s + N0

with

N4  =

0 00 0

,   N3  =

0 00 1

,   N2 =

1 10   −4

,   N1 =

 6 2−6 5

,   and   N0 =

9   −36   −2

.

Therefore, a block controller–form realization (A, B, C, D) of  H(s) is given by

A =

0 I2   0 00 0 I2   00 0 0 I2

−d0I2   −d1I2   −d2I2   −d3I2

 =

0 0 1 0 0 0 0 00 0 0 1 0 0 0 00 0 0 0 1 0 0 00 0 0 0 0 1 0 00 0 0 0 0 0 1 00 0 0 0 0 0 0 1

−9 0 12 0 2 0   −4 00   −9 0 12 0 2 0   −4

,

B =

000I2

 =

0 00 00 00 00 00 01 00 1

,

C = N0   N1   N2   N3

=

9   −3 6 2 1 1 0 06   −2   −6 5 0   −4 0 1

,

and

D =  N4 =

0 00 0

.

Page 117: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 117/119

EE 580 — Linear Control Systems    7

This realization is of order 8; it is controllable but not observable. Reducing (A, B, C, D) to thestandard form for unobservable systems and taking the controllable and observable part, we obtaina controllable and observable realization (Aco, Bco, Cco, Dco) of  H(s):

Aco  = 17/33   −70/33 31/99 280/99   −34/29780/11   −178/11   −320/33 745/33   −160/9919/11   −20/11   −76/33 80/33 61/9948/11   −87/11   −64/11 116/11   −32/33

−21/11   −30/11 28/11 40/11 14/33

,   Bco  = −17/297 70/297−80/99 178/99

−19/99 20/99−16/33 29/33

7/33 10/33

,

Cco =

9   −3 6 2 16   −2   −6 5 0

,   and   Dco =

0 00 0

.

This fifth-order realization is of minimal order. Note that, in this example, the block controller–formrealization is not of minimal order, and that the minimal order is not equal to the degree of  d(s)either. Since (Aco, Bco, Cco, Dco) is both controllable and observable, one can in turn transform itto its controller and observer canonical forms. In general, if 

 H(s) ∈ R(s)l×m and if the degree of the

denominator p olynomial is r, then the minimal order  n satisfies  r ≤ n ≤ mr; see Theorem 13.13.

13.6 Other Realization Algorithms

13.6.1 Diagonal Realizations [2, Section 8.4.3][1, Section 6.1]

Given a strictly proper H(s)   ∈   R(s)l×m, let   d(s) be the monic least common multiple of thedenominators of the entries of  H(s). Suppose  d(s) has distinct real roots  λ1, . . . ,  λr  such that

d(s) = (s − λ1) · · · (s − λr).

Then we can expand H(s) into partial fractions and write

 H(s) = N(s)/d(s) =

ri=1

Ri/(s − λi),

where the residue matrices Ri  can be obtained as

Ri  = lims→λi

(s − λi) H(s).

If  ρi   is the rank of  Ri, then we can perform an LU factorization of  Ri  and write

Ri =  CiBi

for some full-column-rank matrix Ci ∈ Rl×ρi and some full-row-rank matrix  Bi  ∈ R

ρi×m. Then it

is immediate that (A, B, C), with

A =

λ1Iρ1   0. . .

0   λrIρr

∈ R(r

1 ρj)×(

r1 ρj),

B =

B1...

Br

∈ R(r

1 ρj)×m,   C =

C1   · · ·   Cr

∈ R

l×(r

1 ρj),

Page 118: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 118/119

EE 580 — Linear Control Systems    8

is a realization of  H(s). The realization has order r

 j=1 ρ j.Let n =

 r j=1 ρ j. The controllability matrix  C  of the diagonal realization can be written

C =  BdV

where

Bd = B1   0

. . .

0 Br

  and   V = Im   λ1Im   · · ·   λn−11   Im

...   ...   ...Im   λrIm   · · ·   λn−1

r   Im

.

The matrix  V   is a block  Vandermonde matrix , which is of full rank as long as  λi  =  λ j   for   i  =  j.Also, by construction, the matrix  Bd  has full rank, and hence the realization is controllable. Theobservability can be shown similarly. Therefore, the diagonal realization, if exists, is minimal.

13.6.2 Realization via Markov Parameters [3, Section 5.4D]

As in the previous section, let

d(s) = sr + dr−1sr−1 + · · · + d1s + d0

be the denominator matrix of a strictly proper H(s) ∈ R(s)l×

m. The transfer function matrix H(s)can be written  H(s) = H1s−1 + H2s−2 + · · ·  ,

where  Hi  are the Markov parameters of  H(s). Then (A, B, C) given by

A =

0 Il   · · ·   0...

  ...  . . .

  ...0 0   · · ·   Il

−d0Il   −d1Il   · · · −dr−1Il

∈ Rrl×rl,   B =

H1...

Hr

∈ Rrl×m,

C =

Il   0   · · ·   0

∈ R

l×rl

is a realization of  H(s). The proof goes as follows:It is easily seen that

CCA

...CAr−1

 =

Il   0. . .

0 Il

,   (13.12)

and hence that  CAi−1B   =  Hi   for   i  = 1, . . . ,   r. Since H(s) is strictly proper, thereexist R1, . . . ,  Rr  ∈ R

l×m such that

d(s)

 H(s) = (sr + dr−1sr−1 + · · · + d1s + d0)(H1s−1 + H2s−2 + · · · )

= R1sr−1 + · · · + Rr−1s + Rr.

Equating coefficients of equal powers of  s, we obtain

H1 = R1,

H2 = R2 − dr−1H1,

...

Hr = Rr − dr−1Hr−1 − · · · − d1H1,

Hi+r =   − dr−1Hi+r−1 − · · · − d1Hi+1 − d0Hi, i = 1,   2, . . . .

Page 119: ee580_notes.pdf

7/21/2019 ee580_notes.pdf

http://slidepdf.com/reader/full/ee580notespdf 119/119

EE 580 — Linear Control Systems    9

Using this, it can be shown that  CAi−1B =  Hi  for all i  = 1, 2, . . . , which is equivalentto (A, B, C) being a realization of  H(s).

The realization (A, B, C) has order  rl, and in view of (13.12) it is observable. Similarly, dualityarguments lead to a dual-form controllable realization.

13.7 Determination of Minimal Order [2, Section 8.3.3]

Let H(s)  ∈  Rl×m be proper. The determinant of any square submatrix of  H(s) is called a  minor 

of  H(s). The monic least common multiple of the denominators of “all nonzero minors” of  H(s) iscalled a  pole polynomial   of  H(s) (and the roots of the pole polynomial of  H(s) are called the  poles of  H(s)). The degree of the pole polynomial of  H(s) is called the  McMillan degree   of  H(s).

Theorem 13.13  [3, Chapter 5, Theorem 3.11] If  (A, B, C, D) is a minimal realization of a transfer  function matrix  H(s), then the pole polynomial of  H(s)   is equal to the characteristic polynomial of  A, and the McMillan degree of  H(s)  equals the order of  (A, B, C, D).

The  Hankel matrix  of a proper transfer function

 H(s) ∈ R

l×m is defined as

MH (i, j) =

H1   H2   · · ·   H jH2   H3   · · ·   H j+1

...  ...

  ...Hi   Hi+1   · · ·   Hi+ j−1

,

where  H1,   H2, . . . are the Markov parameters such that H(s) =  H0 +  H1s−1 + H2s−2 + · · · . If (A, B, C, D) is any realization of  H(s), then we have  MH (i, j) =  O(i)C( j), where

O(i) =

CT (CA)T · · ·

CAi−1TT ,   C( j) =

B AB   · · ·   A j−1B

.

Theorem 13.14   [3, Chapter 5, Theorem 3.13]   If   r   is the degree of the denominator polynomial (i.e., the least common multiple of the denominators of the entries) of a proper transfer func-

tion  H(s), then the order of a minimal realization of  H(s)  is equal to  rank MH (r, r).

For example, the pole polynomial of the transfer function matrix H(s) in Section 13.5 is

 pH(s) = (s − 1)2(s + 3)3,