Document 10909084

advertisement
Hindawi Publishing Corporation
Advances in Decision Sciences
Volume 2010, Article ID 810605, 19 pages
doi:10.1155/2010/810605
Research Article
Discretizing LTI Descriptor (Regular) Differential
Input Systems with Consistent Initial Conditions
Athanasios D. Karageorgos,1 Athanasios A. Pantelous,2
and Grigoris I. Kalogeropoulos1
1
2
Department of Mathematics, University of Athens, 15784 Athens, Greece
Department of Mathematical Sciences, University of Liverpool, Liverpool L69 3BX, UK
Correspondence should be addressed to Athanasios A. Pantelous, a.pantelous@liverpool.ac.uk
Received 4 April 2009; Accepted 18 January 2010
Academic Editor: Graeme Wake
Copyright q 2010 Athanasios D. Karageorgos et al. This is an open access article distributed under
the Creative Commons Attribution License, which permits unrestricted use, distribution, and
reproduction in any medium, provided the original work is properly cited.
A technique for discretizing efficiently the solution of a Linear descriptor regular differential
input system with consistent initial conditions, and Time-Invariant coefficients LTI is introduced
and fully discussed. Additionally, an upper bound for the error xkT − xk that derives from the
procedure of discretization is also provided. Practically speaking, we are interested in such kind of
systems, since they are inherent in many physical, economical and engineering phenomena.
1. Introduction: Preliminary Results
During the discretization or sampling process, we should replace the original continuoustime systems with finite sequences of values at specified discrete-time points. This important
process is commonly used whenever the differential systems involve digital inputs, and
by having numerical data, the sampling operation and the quantization are necessary.
Additionally, the discretization or sampling process is occurred whenever significant
measurements for the system are obtained in an intermittent fashion. For instance, we can
consider a radar tracking system, where there is information about the azimuth and the
elevation, which is obtained as the antenna of the radar rotates. Consequently, the scanning
operation of the radar produces many important sampled data.
In our approach, we consider the LTI descriptor or generalized differential input
systems of type
Eẋt Axt But,
1.1
2
Advances in Decision Sciences
where matrices E, A ∈ Mn × n, F i.e., M is the algebra of square matrices with elements in
the field F and B ∈ Mn × l, F are constants; the state xt ∈ Mn × 1, F has consistent initial
conditions. We shall call x0 a consistent initial condition for 1.1 at to if there is a solution for
1.1, which is defined on some interval to , to γ , γ > 0 such that xto xo ∈ Mn × 1, F,
the input ut ∈ Ml × 1, F, and the det E 0 are related to the matrix pencil theory, since its
algebraic, geometric, and dynamic properties stem from the structure of the associated pencil,
that is, sE − A. Moreover, for the sake of simplicity, we set in the sequel Mn Mn × n; F and
Mnm Mn × m; F.
Now, in what it follows, the pencil sE − A is regular, that is, detsE − A /
0.
Practically speaking, descriptor or generalized regular or singular differential
systems constitute a more general class than linear state space systems do. Considering
applications, these kinds of systems appear in the modelling procedure of many physical,
engineering, mechanical, actuarial, and financial problems. For instance, in engineering, in
electrical networks, and in constrained mechanics, the reader may consult 1–6 , and so forth.
In Economics, the famous Leontief input-output singular dynamic model is well known; see
for instance some of the numerous references 2, 3, 7–12 , and so forth.
In this paper, we provide two main research directions that are being summarized
briefly below.
i First, we want to provide a computationally efficient method for discretizing
LTI descriptor regular differential systems with input signals and consistent initial
conditions. The consistency of the initial conditions is necessary, because it
eliminates completely the possibility to appear as a distributional expression for
the solution of system 1.1, that is, Dirac delta functions and its derivatives.
ii Second, according to the authors’ knowledge, for the first time an upper bound
for the error xkT − xk , which is derived during the discretization process, is
finally obtained. Consequently, an analytic expression that penalizes our choice
for the sampling period T is provided through the notion of the · -norm of the
difference between the continuous-time solution xkT , at time t kT , and the
relevant discrete-time points xk .
This investigation is relevant to and it extends further the recent work proposed by
Karampetakis; see 13 and Kalogeropoulos et al.; see 14 . Concerning the mathematical
tools, only the Weierstrass canonical form WCF and some fundamental elements of matrix
pencil theory are required.
Recently, in 15, 16 , several numerical issues of the WCF of a regular matrix pencil
are presented and discussed. Thus, briefly speaking, in these research papers, two important
computational tools are considered: a the QZ algorithm to specify the required root range
of the pencil and b the updating technique to compute the index of annihilation; see 16 .
The proposed updating technique takes advantages of the already computed rank of the
sequences of matrices that appears during our procedure reducing significantly the required
floating-point operations. The algorithms are implemented in a numerical stable manner,
giving efficient results.
For reasons of convenience, some basic concepts and definitions from matrix pencil
theory are introduced; see for more details 2–4, 17–22 et al. Thus, the class of strict
equivalence is characterized by a uniquely defined element, known as a WCF, that is,
sEw − Aw . Consequently, when the pencil sE − A is regular, we have elementary divisors
of the following type:
Advances in Decision Sciences
3
i e.d. of the type sd , d ∈ N, are called zero finite elementary divisors z. f.e.d.;
0, c ∈ N are called nonzero finite elementary divisors nz.
ii e.d. of the type s − ac , a /
f.e.d.;
iii e.d. of the type sq are called infinite elementary divisors i.e.d..
Then, the Weierstrass canonical form of the regular pencil sE − A is defined by
sEw − Aw block diag sIp − Jp , sHq − Iq ,
1.2
where the first normal Jordan type block sIp − Jp is uniquely defined by the set of f.e.d.
ν
pj p
s − a1 p1 , . . . , s − aν pν ,
1.3
j1
of sF − G and has the form
sIp − Jp block diag sIp1 − Jp1 a1 , . . . , sIpν − Jpν aν .
1.4
And the q blocks of the second uniquely defined block sHq − Iq correspond to the i.e.d.
σ
sqσ ,
sq1 , . . . , qj q
1.5
j1
of sE − A and has the form
sHq − Iq block diag sHq1 − Iq1 , . . . , sHqσ − Iqσ .
1.6
q∗
Thus the Hq is a nilpotent matrix of index q∗ max{qj : j 1, 2, . . . , σ}, where Hq O
and Ipi , Jpi ai , Hqi are the matrices
⎡
1
⎢
⎢0
⎢
Ipi ⎢
⎢ ..
⎢.
⎣
0
0 ···
1 ···
.. . .
.
.
0 ···
⎡
0
⎢
⎢0
⎢
⎢
⎢
Hqi ⎢ ...
⎢
⎢
⎢0
⎣
⎤
0
⎥
0⎥
⎥
⎥ ∈ M p i × pi , R ,
.. ⎥
.⎥
⎦
1
⎡
ai
⎢
⎢0
⎢
⎢
⎢
Jpi ai ⎢ ...
⎢
⎢
⎢0
⎣
1
⎤
1 0 ··· 0
⎥
0 1 · · · 0⎥
⎥
⎥
.. . . .. .. ⎥ ∈ M q × q , R ,
i
i
. . .⎥
.
⎥
⎥
0 0 0 1⎥
⎦
0 0 0
0 ··· 0
0 0
⎤
⎥
ai 1 · · · 0 ⎥
⎥
⎥
.. . . .. .. ⎥ ∈ M p × p , C .
i
i
. . .⎥
.
⎥
⎥
0 0 ai 1 ⎥
⎦
0 0
0
0 ai
1.7
4
Advances in Decision Sciences
Now, considering 4, 17, 18 , and the transformation
xt Qyt,
1.8
we obtain the following result. Note that matrix Q is not unique. However, different matrices
Q are mutually related. We use the above transformation, 1.8, since we want to create the
WCF; see 17 .
Theorem 1.1 see 4, 23 . System 1.1 has the following solution:
xt Qn,p e
where Q Qn,p
Jp t−to y to e
p
Qn,q , B t
Jp t−s
Bp,l usds − Qn,q
to
∗
q
−1
i
1.9
Hqi Bq,l u t,
i0
y to Bp,l p
,
and
yt
Q−1 xto .
o
B
y t q,l
o
q
However, 1.9 should be transposed to 1.12, which is practically more useful. Thus,
we obtain
xt Qn,p e
Jp t−to y to Qn,q y to Qn,p
p
q
− Qn,q y to − Qn,q
∗
−1
q
q
Qn,p
t
eJp t−s Bp,l usds
to
i
Hqi Bq,l u t
i0
⎤
⎡
t
eJp t−to Op,q yp to ⎦ Qn,p eJp t−s Bp,l usds
⎣
Qn,q
y to Oq,p
Iq
to
1.10
q
Qn,q −y to −
q
∗
q
−1
i
Hqi Bq,l u t
.
i0
In order the system 1.1 to obtain consistent initial conditions, we should consider that
⎡
⎤
y to ⎣ p
⎦ Q−1 xto ,
y to q
−y to ∗
q
−1
q
i
Hqi Bq,l u to .
1.11
i0
Consequently, we obtain the desired expression for 1.9, that is,
xt Q
eJp t−to Op,q
Oq,p
Qn,q
∗
−1
q
i0
Iq
Q−1 xto Qn,p
t
to
i
Hqi Bq,l ui to − u t .
eJp t−s Bp,l usds
1.12
Advances in Decision Sciences
5
Moreover, by definition, the state-transition matrix of the autonomous linear
descriptor differential system, that is, Eẋt Axt, is given by
Φt, to Q
eJp t−to Op,q
Oq,p
Iq
Q−1 .
1.13
Finally, after some simple algebraic calculations, we obtain another more elegance
form for the solution of system 1.1, that is,
xt Φt, to xto t
Φt, sQn,p Bp,l usds Qn,q
∗
−1
q
to
i
Hqi Bq,l ui to − u t ,
1.14
i0
where xto ∈ T belongs to the space of consistent initial conditions of system 1.1, that is,
T
x ∈ Mn × 1, C : x Qn,p Cp − Qn,q
∗
−1
q
i
Hqi Bq,l u to ,
i0
with Cp ∈ M p × 1, C
.
1.15
2. Discretisation of Nonhomogeneous LTI Descriptor Regular
Differential System
In this section, we provide a computationally efficient method for the analytic discretization
of LTI descriptor regular differential systems with input signal and consistent initial
conditions. In the vast literature of descriptor systems, Karampetakis has proposed two
discretization methods for regular systems without see 13, Section 3 and with inputs see
13, Section 4 , which are based on the research work by Koumboulis and Mertzios, see
24 , concerning the solution of regular systems in terms of the Laurent expansion terms of
sE − A−1 . The methodology of both papers, that is, 13 and the present, is equivalent, since
they use the zero-order hold approximation.
Moreover, specifically for regular systems, Rachid, see 21 , has proposed a different
method, which discretizes the exact solution using Euler approximation techniques.
However, in the present paper, a different to the above research works discretization
technique for LTI descriptor differential input systems is presented, which extends further the
research work 8, 14 .
First, we denote T > 0 as the constant sampling period. Without loss of generality and
because of the complex notations that follow, we assume that to 0, and the input function
ut changes only through the time moments tk kT for k 1, 2, . . ., that is,
ut ukT uk ,
2.1
for every t ∈ kT, k 1T .
In more details, 2.1 implies that the input ut is being inserted into a mechanism of
zero-order change after the end of each sampling process. Thus, the input remains constant
from time tk kT until the next sampling period at time tk1 k 1T takes place.
6
Advances in Decision Sciences
Thus, the solution 1.14 is given by
xt Φt, 0x0 t
Φt, sQn,p Bp,l usds Qn,q
∗
−1
q
0
i
Hqi Bq,l ui 0 − u t ,
2.2
i0
where x0 x0 ∈ T.
Moreover, hereafter, we use the notation
xk ≡ xkT ∀k 0, 1, 2, . . . .
2.3
The following theorem provides us with an analytic formula based on 2.2.
Theorem 2.1. An analytic formula for the discretized solution 2.2 of system 1.1 is given by 2.4
xk ΦkT, 0x0 Qn,q
∗
q
−1
k−1 T
0
j0
Hqi Bq,l
Φ jT λ, 0 dλ · Qn,p Bp,l uk−1−j
ui 0 − u kT .
2.4
i
i0
Proof. First, we consider 2.2 at time moments t kT and t k 1T .
Thus, we obtain for the time t kT
xk ΦkT, 0x0 kT
ΦkT, sQn,p Bp,l usds Qn,q
0
∗
−1
q
i
Hqi Bq,l ui 0 − u kT ,
2.5
i0
and for the time t k 1T
xk1 Φk 1T, 0x0 Qn,q
∗
−1
q
Hqi Bq,l
k1T
Φk 1T, sQn,p Bp,l usds
0
ui 0 − u k 1T .
2.6
i
i0
The state-transition matrix 1.13 satisfies the following two properties:
ΦT, 0 · ΦkT, 0 Φk 1T, 0,
2.7
ΦT, 0 · ΦkT, s Φk 1T, s.
2.8
Advances in Decision Sciences
7
Now, we multiply 2.5 from the left with ΦT, 0, and we consider 2.7 and 2.8. Thus, we
obtain
ΦT, 0xk Φk 1T, 0x0 ΦT, 0Qn,q
∗
−1
q
kT
Φk 1T, sQn,p Bp,l usds
0
Hqi Bq,l
2.9
u 0 − u kT .
i
i
i0
Consequently, by also using 2.6 and 2.9, we take the difference
xk1 − ΦT, 0xk k1T
Φk 1T, sQn,p Bp,l usds
kT
Qn,q
∗
−1
q
Hqi Bq,l ui 0 − ui k 1T 2.10
i0
− ΦT, 0Qn,q
∗
−1
q
i
Hqi Bq,l ui 0 − u kT .
i0
Afterwards, we should define
Q
−1
⎤
⎡
p,n
Q
⎦,
⎣
p,n
Q
⎤
⎡
⎤
⎡
⎡
p,n p,n
p,n Qn,p
Q
Q
Q
⎦Q ⎣
⎦ Qn,p Qn,q ⎣
Q−1 Q ⎣
q,n
q,n
q,n Qn,p
Q
Q
Q
p,n Qn,q
Q
q,n Qn,q
Q
⎤
2.11
⎦ In ,
with
p,n Qn,q Op,q ,
Q
p,n Qn,p Ip ,
Q
q,n Qn,p Oq,p ,
Q
2.12
q,n Qn,q Iq .
Q
Thus,
ΦT, 0Qn,q Q
eJp T Op,q
Oq,p
Iq
Qn,p Qn,q
Q−1 Qn,q Qn,p
eJp T Op,q Op,q
Oq,p
Iq
Iq
⎤
⎡
pn Qnq
eJp T Op,q Q
⎦
⎣
Qn,q
qn Qnq
Oq,p Iq
Q
Qn,p , Qn,q
Op,q
Iq
Qn,q .
2.13
8
Advances in Decision Sciences
So ΦT, 0Qn,q Qn,q . Consequently, 2.10 is transformed to 2.14
xk1 ΦT, 0xk Qn,q
∗
−1
q
k1T
Φk 1T, sQn,p Bp,l usds
kT
i
2.14
Hqi Bq,l ui kT − u k 1T .
i0
Note that by denoting s kT w, the integral
k1T
Φk 1T, sQn,p Bp,l usds T
kT
k1T
kT
Φk 1T, sQn,p Bp,l usds is given by
Φk 1T, kT wQn,p Bp,l ukT wdw
0
T
2.1
0
2.15
Φk 1T, kT wdw · Qn,p Bp,l uk .
Moreover, by using 2.7 and 2.8, we obtain
T
Φk 1T, kT wdw T
0
Φk 1T, kT Φ0, wdw.
2.16
0
And considering again 2.8, we have
Φk 1T, kT ΦT, 0 · ΦkT, kT ΦT, 0
2.17
because ΦkT, kT In . Finally, we obtain
T
Φk 1T, kT wdw 0
T
ΦT, 0Φ0, wdw 0
T
ΦT − w, 0dw T
0
T
ΦT, wdw
0
2.18
Φλ, 0dλ,
0
when we define λ T − w. Thus,
k1T
Φk 1T, sQn,p Bp,l usds kT
T
0
Φλ, 0dλ · Qn,p Bp,l uk .
2.19
Combining 2.14 and 2.19, we take the recursive formula 2.20
xk1 ΦT, 0xk Qn,q
∗
−1
q
i0
T
0
Φλ, 0dλ · Qn,p Bp,l uk
Hqi Bq,l
u kT − u k 1T .
i
i
2.20
Advances in Decision Sciences
9
Then, the formula 2.4 is obtained by 2.20. Analytically,
i for k 0, we take
x1 ΦT, 0x0 T
0
i
Hqi Bq,l ui 0 − u T ;
∗
−1
q
Φλ, 0dλ · Qn,p Bp,l u0 Qn,q
2.21
i0
ii for k 1, we have
T
x2 ΦT, 0x1 0
Φλ, 0dλ · Qn,p Bp,l u1 Qn,q
i
Hqi Bq,l ui T − u 2T i0
ΦT, 0ΦT, 0x0 ΦT, 0Qn,q
∗
−1
q
∗
−1
q
T
0
ΦT, 0Φλ, 0dλ · Qn,p Bp,l u0
Hqi Bq,l
2.22
u 0 − u T i
i
i0
T
0
Φλ, 0dλ · Qn,p Bp,l u1 Qn,q
∗
−1
q
i
Hqi Bq,l ui T − u 2T .
i0
Since ΦT, 0Qn,q Qn,q , we conclude to
x2 Φ2T, 0x0 T
0
T
0
ΦT λ, 0dλ · Qn,p Bp,l u0
Φλ, 0dλ · Qnp Bp,l u1 Qn,q
∗
−1
q
i
2.23
Hqi Bq,l ui 0 − u 2T ;
i0
iii for k 3, we take
x3 Φ3T, 0x0 T
0
T
0
T
Φ2T λ, 0dλ · Qn,p Bp,l u0 Φλ, 0dλ · Qn,p Bp,l u2 Qn,q
∗
−1
q
Hqi Bq,l
0
ΦT λ, 0dλ · Qn,p · Bp,l u1
i
i
2.24
u 0 − u 3T .
i0
Continuing as above, we finally obtain the analytic formula 2.4.
Now, in order to complete the discretization process of the solution 2.2 of the system
i
i
1.1, we should replace the ith-order derivative uk ≡ u kT with expressions of ukT .
Thus, the following result is used, see 25 ,
i
i
uk u kT ≈
i
j0
−1j
i
Ti
j
uk−ij
.
2.25
10
Advances in Decision Sciences
Profoundly, based on 2.25, the following difference can be easily proved.
i1
i
i
uk − uk1 ≈
j1
j0 −1
i1
j
uk−ij
T i1
2.26
.
Remark 2.2. The discrete-time solution which represents 2.2 of the continuous-time system
1.1 is given by 2.27, see also 13, Theorem 3 xk ΦkT, 0x0 Qn,q
k−1 T
j0
∗
−1
q
1
i0
T i1
0
Φ jT λ, 0 dλ · Qn,p Bp,l uk−1−j
Hqi Bq,l
i1
2.27
−1j1
j0
i1
j
uk−1j .
Remark 2.3. Equation 2.20 reminds us with a new discrete-time system of the form
xk1 Axk Bσuk ,
2.28
i
where A ΦT, 0, whereas the differences ui kT − u k 1T can be replaced by 2.26
in order to get the matrix Bσ. This approach is equivalent to Theorem 3.7 in 13 .
Remark 2.4. The induction method show that it is true for n 0; assume that it is true for
n k; try to prove that it is true for n k 1 can prove 2.4. However, we strongly believe
that our approach seems to be more natural, since 2.4 is analytically constructed.
In the next section, according to the authors’ knowledge, an upper bound for the error
xkT − xk , which is derived during the discretization process, is provided for the first time.
3. Error Analysis for the Discretization (or Sampling) Process
In this section, we provide an analytic expression that penalizes our choice for the sampling
period T through the notion of the Euclidean · -norm of the difference between the
continuous-time solution 1.13, at time t kT , xkT and the relevant discrete-time points
xk , see 2.4, that is, we are interested in the difference xkT − xk .
Following the results of the previous sections, we obtain 3.2. Analytically, we firstly
note that the comparison of the solutions is provided at time moment t kT .
The solution 1.13 at time moment t kT for k 1, 2, 3, . . . , is given by 3.1
xkT ΦkT, 0x0 Qn,q
∗
−1
q
i0
kT
ΦkT, sQn,p Bp,l usds
0
Hqi Bq,l ui 0 − ui kT 3.1
Advances in Decision Sciences
11
Combining 2.4 and 3.1, we obtain the following important expression:
xkT − xk kT
ΦkT, sQn,p Bp,l usds
0
−
k−1 T
j0
0
3.2
Φ jT λ, 0 dλ · Qn,p Bp,l uk−1−j
note that it is not important to consider the completely discretized equation 2.27 instead of
2.4, since the second part with the ith-order derivatives is eliminated.
In order to compare the two solutions at a fixed time moments, the following lemmas
are stated and straightforwardly proved.
Lemma 3.1. The equality 3.3 holds
kT
ΦkT, sQn,p Bp,l usds k−1 j1T
0
j0
jT
ΦkT, sds · Qn,p Bp,l uj .
3.3
Proof. Consider that
kT
ΦkT, sQn,p Bp,l usds 0
T
ΦkT, sQn,p Bp,l usds 2T
0
··· ΦkT, sQn,p Bp,l usds
T
kT
k−1T
3.4
ΦkT, sQn,p Bp,l usds.
Since 2.1 holds, we take
kT
ΦkT, sQn,p Bp,l usds 0
T
0
ΦkT, sQn,p Bp,l u0 ds ··· T
0
k−1T
kT
k−1T
T
ΦkT, sQn,p Bp,l u1 ds
ΦkT, sQn,p Bp,l uk−1 ds
ΦkT, sds · Qn,p Bp,l u0 ··· Thus, 3.3 derives.
kT
2T
2T
T
3.5
ΦkT, sds · Qn,p Bp,l u1
ΦkT, sds · Qn,p Bp,l uk−1 .
12
Advances in Decision Sciences
Lemma 3.2. The equality 3.6 holds
j1T
T
ΦkT, sds jT
Φ k − j T − λ, 0 dλ.
3.6
0
Proof. We define that s jT λ; if s jT , then λ 0; if s j 1T, then λ T .
Thus, we obtain
j1T
ΦkT, sds jT
T
Φ kT, jT λ dλ 0
T
Φ k − j T − λ, 0 dλ
3.7
0
since it can easily be proven that
Φ kT, jT λ Φ k − j T − λ, 0 .
3.8
Now, we consider Lemmas 3.1 and 3.2, and 3.2 is transformed to 3.9
xkT − xk k−1 T
j0
−
0
Φ k − j T − λ, 0 dλ · Qn,p Bp,l uj
k−1 T
j0
0
3.9
Φ jT λ, 0 dλ · Qn,p Bp,l uk−1−j .
Lemma 3.3. The equality 3.10 holds
k−1 T
j0
0
Φ jT λ, 0 dλ · Qn,p Bp,l uk−1−j k−1 T
j0
0
Φ k − 1 − j T λ, 0 dλ · Qn,p Bp,l uj .
3.10
Proof. Making the transformation k − 1 − j i, we obtain j k − 1 − i.
Since j 0, 1, . . . , k − 1, then i k − 1, k − 2, . . . , 0. So, we obtain
k−1 T
j0
0
Φ jT λ, 0 dλ · Qn,p Bp,l uk−1−j k−1
i0
Now, if we also define that j i, 3.10 holds.
Φk − 1 − iT λ, 0dλ · Qn,p Bp,l ui .
3.11
Advances in Decision Sciences
13
Following the results of Lemma 3.3, 3.9 is obtained the following expression:
xkT − xk k−1 T
Φ k − j T − λ, 0 − Φ k − 1 − j T λ, 0 dλ · Qn,p Bp,l uj .
0
j0
3.12
Moreover, it is true that
Φ k − j T − λ, 0 − Φ k − j T − T λ, 0 Φ k − j T, 0 Φ−λ, 0 − Φλ − T, 0.
3.13
Consequently, 3.12 is given by
xkT − xk k−1
Φ k − j T, 0 ·
T
0
j0
Φ−λ, 0 − Φλ − T, 0dλ · Qn,p Bp,l uj ,
3.14
which is practically more significant in what follows.
For the calculation of the upper bound of the difference xkT −xk , we use the Euclidean
· -norm for vectors. Thus, we take
xkT − xk ≤
k−1
Φ k − j T, 0 ·
T
0
j0
k−1
·
Φ k − j T, 0
Φ−λ, 0 − Φλ − T, 0dλ · Qn,p Bp,l uj
T
0
j0
Φ−λ, 0 − Φλ − T, 0dλ Qn,p · Bp,l · uj .
3.15
Now, for the calculation of the upper bound of the difference 3.15, one more lemma
and some propositions should be considered.
Lemma 3.4. The inequality 3.16 holds
e
Jp t
≤
ν
i1
⎡
⎣e
αi t
pi −1! "2
tm
m0
m!
1/2 ⎤
⎦.
pi − m
3.16
Proof. Considering 1.7, the block matrix Jp is given by
⎤
⎡
Jp1 α1 ⎢
⎢
Jp2 α2 ⎢
Jp ⎢
⎢
..
⎢
.
⎣
⎥
⎥
⎥
⎥,
⎥
⎥
⎦
Jpν αν 3.17
14
Advances in Decision Sciences
where α1 , α2 , . . . , αν are nonzero finite elementary divisors of matrix pencil sE − A; consult
Section 1 for more details. Moreover,
⎤
⎡
αi 1
⎥
⎢
⎥
⎢
⎢ αi . . . ⎥
⎥
⎢
Jpi αi ⎢
⎥
⎢
.
.. . ⎥
⎥
⎢
.
.
⎦
⎣
αi p ×p
i
for every i 1, 2, . . . , v.
3.18
i
Then
e Jp t
⎤
⎡ Jp α1 t
e 1
⎢
⎢
eJp2 α2 t
⎢
⎢
⎢
..
⎢
.
⎣
⎥
⎥
⎥
⎥
⎥
⎥
⎦
with p1 p2 · · · pν p.
3.19
eJpν αν t
Considering the well-known property
√
√
√
x y ≤ x y, it is derived that
eJp t ≤ eJp1 α1 t · · · eJpν αν t ν
eJpi αi t .
3.20
i1
Note that
eJpi αi t
⎡
t2
1
t
···
⎢
2!
⎢
⎢
⎢
⎢
⎢
⎢0 1 t · · ·
⎢
⎢
⎢
eαi t ⎢
⎢
.. ..
⎢
. .
⎢
⎢
⎢
⎢
..
⎢
.
⎢
⎣
⎤
tpi −1
⎥
pi − 1 ! ⎥
⎥
⎥
⎥
tpi −2 ⎥
⎥
pi − 2 ! ⎥
⎥
⎥
⎥.
⎥
..
⎥
.
⎥
⎥
⎥
⎥
⎥
t
⎥
⎦
1
3.21
Advances in Decision Sciences
15
Consequently, by using some basic properties of the Euclidean norm, we take
⎡
e
Jpi αi t
⎣ e
αi t 2
pi te
αi t 2
pi − 1 · · · pi − pi − 1
e
p
i −1
2αi t
!
m0
tm
m!
"2
t2 αi t
e
2!
tpi −1
eαi t
pi − 1!
2
pi − 2
2 ⎤1/2
⎦
1/2
pi − m
e
αi t
3.22
pi −1! "2
tm
m!
m0
1/2
pi − m
.
Thus, we conclude to the following result:
e
Jpi αi t
e
αi t
pi −1 ! "2
tm
m0
m!
1/2
pi − m
3.23
,
for every i 1, 2, . . . , ν. Hence, inequality 3.16 holds.
Proposition 3.5. The inequality 3.24 holds
Φ k − j T, 0
≤ Q · Q−1
⎛⎛
⎞
pi −1 !
1/2 ⎞k−j
"
2
ν
m
⎜⎝ αi T T
⎠ (q⎟
e
pi − m
⎝
⎠.
m!
m0
i1
3.24
Proof. We known that
Jp k−jT
Op,q −1
e
Q .
Φ k − j T, 0 Q
Oq,p
Iq
3.25
Then, using also the well-known property of the Euclidean norm
obtain
Φ k − j T, 0
*
xy ≤
≤ Q eJp k−jT Iq
Q−1
Q · Q−1
since we have Iq √
12 12 , · · · 1.2 +
√
eJp k−jT ( q ,
√
x
√
y, we
3.26
q.
q
Moreover,
eJp k−jT eJp T eJp T · · · eJp T ≤
e Jp T
k−j
.
3.27
16
Advances in Decision Sciences
Considering, now, Lemma 3.4, we have
eJp κ−jT
⎛ ⎡
pi −1!
1/2 ⎤⎞k−j
ν
m "2
T
⎦⎠ .
≤ ⎝ ⎣eαi T
pi − m
m!
m0
i1
3.28
Thus, the statement of Proposition 3.5 derives.
Proposition 3.6. The inequality 3.29 holds
Φ−λ, 0 − Φλ − T, 0
≤ Q · Q−1
⎧
⎡
pi −1!
1/2
ν
⎨
−1m λm "2
−α
λ
i
⎣e
·
pi − m
⎩ i1
m!
m0
eαi λ−T pi −1!
λ − T m "2
m!
m0
3.29
1/2 ⎤⎫
⎬
⎦
pi − m
⎭
Proof. It is known that
Φ−λ, 0 − Φλ − T, 0 Q
Oq,p
Q
e−Jp λ Op,q
Iq
Q
−1
−Q
e−Jp λ − eJp λ−T Op,q
Oq,p
Op
Jp λ−T Op,q
e
Oq,p
Iq
Q−1
3.30
Q−1 .
Then, the norm is
Φ−λ, 0 − Φλ − T, 0 ≤ Q · e−Jp λ − eJp λ−T · Q−1
≤ Q · Q−1
e−Jp λ eJp λ−T 3.31
.
Considering also Lemma 3.4, 3.29 derives.
Now, the whole discussion of this section is completed with the statement of
Theorem 3.7 whose proof is based on Lemmas 3.1 and 3.2 and Propositions 3.5 and 3.6.
Advances in Decision Sciences
17
Theorem 3.7. The upper bound for the error xkT − xk , which is derived during the discretization
process, is given by
xkT − xk
⎫
⎡⎧⎛ ⎡
pi −1!
1/2 ⎤⎞k−j
⎪
⎪
"
2
ν
⎨ m
( ⎬
⎢ ⎝ ⎣ αi T T
⎠
⎦
≤
e
pi − m
q
⎣
⎪
⎪ i1
m!
⎭
m0
j0 ⎩
k−1
⎧ ⎡
1/2
pi −1 !
T ⎨
ν
−1m · λm "2
−α
λ
⎣e i
·
pi − m
m!
0 ⎩ i1
m0
eαi λ−T pi −1 !
λ − T m "2
m0
· Q2 · Q−1
2
m!
⎤
1/2 ⎤⎫
⎬
⎦ dλ · u ⎦
pi − m
j
⎭
· Qn,p · Bp,l .
3.32
Now, with the upper bound derived by 3.32, we have an analytic expression that
penalizes our choice for the sampling period T . Although, the above expression is very
complicated with a lot of parameters involved, it provides an insightful overview of the error
obtained during the discretization or sampling process. The following Remark evaluates
our upper bound. According to the existing literature, see for instance 14 and references
therein, when the sampling period T tends to zero, we are waiting to obtain xkT −xk → 0;
that is, the continuous-time solution at time t kT is exactly the same with the discrete-time
points xk .
Remark 3.8. Consider that the sampling period T tends to zero, that is, T → 0. In the integral,
the function consists exponential and polynomial, continuous functions with respect to λ.
Thus, it is well known from the basic Calculus that
x
lim
x→0 0
ftdt lim Fx − F0 lim Fx − F0 F0 − F0 0.
x→0
x→0
3.33
Moreover, in our case since T → 0, we obtain
⎧ ⎡
pi −1 !
1/2
T⎨
ν
−1m ·λm "2
−α
λ
i
⎣e
lim
pi −m
T → 0 0⎩
m!
m0
i1
pi −1 !
1/2 ⎤⎫
⎬
λ−T m "2
⎦ dλ 0.
eαi λ−T pi −m
⎭
m!
m0
3.34
18
Advances in Decision Sciences
Then
⎫
⎡⎧⎛ ⎡
pi −1!
1/2 ⎤⎞k−j
⎪
ν
k−1 ⎪
⎨ m "2
( ⎬
T
⎢ ⎝ ⎣ αi T
⎠
⎦
e
p
−
m
q
lim
⎣
i
⎪
⎪ i1
T →0
m!
⎭
m0
j0 ⎩
⎧ ⎡
pi −1 !
1/2
T ⎨
ν
−1m · λm "2
−αi λ
⎣
·
pi − m
e
m!
0 ⎩ i1
m0
⎤
⎤⎫
⎪
⎪
pi −1 !
1/2
⎥
⎥⎪
⎬
λ − T m "2
⎥
⎥
eαi λ−T pi − m
⎥ dλ · uj ⎥
⎪
⎦
⎦
m!
⎪
m0
⎪
⎭
· Q2 · Q−1
2
3.35
· Qn,p · Bp,l 0
since ν, k, pi for every i 1, . . . , ν are finite.
Consequently, we have seen that if the sampling period T tends to zero, that is, T →
0, then the solution of the discretized system 1.1, that is, 2.4 or 2.27, is an efficient
approximation of the relative solution of continuous-time system, that is, 1.13.
4. Conclusion
In this paper, we have presented and fully discussed a technique for discretizing efficiently
the solution of a linear descriptor regular differential input system with consistent
initial conditions, and time-invariant coefficients. Additionally, according to the authors’
knowledge, for the first time an upper bound for the error xkT − xk that derives from
the procedure of discretization is provided. This error expression penalizes our choice for
the sampling period T through the notion of the · -norm of the difference between the
continuous-time solution xkT , at time t kT , and the relevant discrete-time points xk .
Finally, the results of this paper can be further enriched by the analytic determination
of an appropriate sampling period interval, that is, T ∈ a, b , which is a future plan
of research. An optimal upper bound is also under consideration. Moreover, we are also
interested in investigating the case that the non-consistent initial conditions are considered,
since distributional solutions are derived.
Acknowledgments
The project is cofunded by the European Social Fund and National Resources—THALIS I. The
authors are very grateful to the anonymous referees for their comments, which improved the
quality of the paper.
Advances in Decision Sciences
19
References
1 K. E. Brenan, S. L. Campbell, and L. R. Petzold, Numerical Solution of Initial-Value Problems in
Differential-Algebraic Equations, vol. 14 of Classics in Applied Mathematics, SIAM, Philadelphia, Pa, USA,
1996.
2 S. L. Campbell, Singular Systems of Differential Equations, vol. 1, Pitman, San Francisco, Calif, USA,
1980.
3 S. L. Campbell, Singular Systems of Differential Equations, vol. 2, Pitman, San Francisco, Calif, USA,
1982.
4 L. Dai, Singular Control Systems, vol. 118 of Lecture Notes in Control and Information Sciences, Springer,
Berlin, Germany, 1989.
5 F. L. Lewis, “A survey of linear singular systems,” Circuits, Systems, and Signal Processing, vol. 5, no.
1, pp. 3–36, 1986.
6 V. L. Mehrmann, The Autonomous Linear Quadratic Control Problem: Theory and Numerical Solution, vol.
163 of Lecture Notes in Control and Information Sciences, Springer, Berlin, Germany, 1991.
7 D. A. Kendrick, “On the Leontief dynamic inverse,” Quarterly Journal of Economics, vol. 86, pp. 693–
696, 1972.
8 R. G. Kreijger and H. Neudecker, “Kendrick’s forward integration method and the dynamic Leontief
multisectoral model,” Quarterly Journal of Economics, vol. 90, pp. 505–507, 1976.
9 W. Leontief, “Mathematics in economics,” Bulletin of the American Mathematical Society, vol. 60, pp.
215–233, 1954.
10 W. Leontief, Essays in Economics, M. E. Sharpe, Inc., New York, NY, USA, 1977.
11 W. Leontief, Input-Output Economics, Oxford University Press, New York, NY, USA, 2nd edition, 1986.
12 D. A. Livesey, “The singularity problem in the dynamic input-output model,” International Journal of
Systems Science, vol. 4, pp. 437–440, 1973.
13 N. P. Karampetakis, “On the discretization of singular systems,” IMA Journal of Mathematical Control
and Information, vol. 21, no. 2, pp. 223–242, 2004.
14 G. I. Kalogeropoulos, A. D. Karageorgos, and A. A. Pantelous, “Discretising effectively a linear
singular differential system by choosing an appropriate sampling period,” IET Control Theory &
Applications, vol. 3, no. 7, pp. 823–833, 2009.
15 G. Kalogeropoulos, M. Mitrouli, A. Pantelous, and D. Triantafyllou, “The weierstrass canonical form
of a regular matrix pencil: numerical issues and computational techniques,” in Proceedings of the 4th
International Conference on Numerical Analysis and Its Applications (NAA ’09), vol. 5434 of Lecture Notes
in Computer Science, pp. 322–329, Springer, Lozenetz, Bulgaria, June 2009.
16 A. D. Karageorgos, On the study of different qualitative characteristics for generalized control systems, Ph.D.
thesis, Athens, Greece, 2009.
17 R. F. Gantmacher, The Theory of Matrices, Chelsea, New York, NY, USA, 1959.
18 G. I. Kalogeropoulos, Matrix pencils and linear systems, Ph.D. thesis, City University, London, UK, 1980.
19 N. Karcanias, “Matrix pencil approach to geometric system theory,” Proceedings of the IEE, vol. 126,
no. 6, pp. 585–590, 1979.
20 N. Karcanias and G. E. Hayton, “Generalised autonomous differential sys-tems, algebraic duality, and
geometric theory,” in Proceedings of the 8th IFAC Triennial World Congress, Kyoto, Japan, 1981.
21 A. Rachid, “A remark on the discretization of singular systems,” Automatica, vol. 31, no. 2, pp. 347–
348, 1995.
22 P. Van Dooren, Reducing Subspaces: Definitions, Properties, and Algorithms, in Matrix Pencils, vol. 973 of
Lecture Notes in Mathematics, Springer, Berlin, Germany, 1983.
23 E. Grispos, Study of generalized linear differential systems with the use of matrix Pencil theory, Ph.D. thesis,
Department of Mathematics, Athens University, Athens, Greece, 1991.
24 F. N. Koumboulis and B. G. Mertzios, “On Kalman’s controllability and observability criteria for
singular systems,” Circuits, Systems, and Signal Processing, vol. 18, no. 3, pp. 269–290, 1999.
25 D. G. Luenberger and A. Arbel, “Singular dynamic Leontief systems,” Econometrica, vol. 45, pp. 991–
995, 1977.
Download