Applied Mathematics and Computation 187 (2007) 250–265 www.elsevier.com/locate/amc
Kronecker operational matrices for fractional calculus and some applications Adem Kilicman *, Zeyad Abdel Aziz Al Zhour Department of Mathematics, University Putra Malaysia (UPM), 43400 Serdang, Selangor, Malaysia Institute for Mathematical Research, University Putra Malaysia (UPM), 43400 Serdang, Selangor, Malaysia Institute of Advanced Technology, University Putra Malaysia (UPM), 43400 Serdang, Selangor, Malaysia
Dedicated to Professor H.M. Srivastava on the occasion of his 65th birthday
Abstract The problems of systems identification, analysis and optimal control have been recently studied using orthogonal functions. The specific orthogonal functions used up to now are the Walsh, the block-pulse, the Laguerre, the Legendre, Haar and many other functions. In the present paper, several operational matrices for integration and differentiation are studied. we introduce the Kronecker convolution product and expanded to the Riemann–Liouville fractional integral of matrices. For some applications, it is often not necessary to compute exact solutions, approximate solutions are sufficient because sometimes computational efforts rapidly increase with the size of matrix functions. Our method is extended to find the exact and approximate solutions of the general system matrix convolution differential equations, the way exists which transform the coupled matrix differential equations into forms for which solutions may be readily computed. Finally, several systems are solved by the new and other approaches and illustrative examples are also considered. 2006 Elsevier Inc. All rights reserved. Keywords: Kronecker product; Convolution product; Kronecker convolution product; Vector operator; Operational matrix; Laplace transform
1. Introduction One of the principle reasons is that matrices arise naturally in communication systems, economic planning, statistics, control theory and other fields of pure and applied mathematics [2,4,7]. Another is because an m · n matrix can be valued as representing a linear map from an n-dimensional vector space to an m-dimensional vector space, and conversely, all such maps can be represented as m · n matrices. The operational matrices for integration and differentiation are studied by many authors [1,3,6,10]. For example, Mouroutsos and
* Corresponding author. Address: Institute of Advanced Technology, University Putra Malaysia (UPM), 43400 Serdang, Selangor, Malaysia. E-mail addresses:
[email protected] (A. Kilicman),
[email protected] (Z.A.A. Al Zhour).
0096-3003/$ - see front matter 2006 Elsevier Inc. All rights reserved. doi:10.1016/j.amc.2006.08.122
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
251
Sparis [6] solved the control problem using Taylor series operational matrices for integration, Chen et al. [3] solved the distributed systems by using Walsh operational matrices and Wang [10] introduced the inversions of rational and irrational transfer functions by using the generalized block pulse operational matrices for differentiation and integration. The Kronecker convolution product is an interesting area of current research and in fact plays an important role in applications, for example, Sumita [9], established the matrix Laguerre transform to calculate matrix convolutions and evaluated a matrix renewal function. In the present paper, several operational matrices for integration and differentiation are studied. we introduce the Kronecker convolution product and expanded to the Riemann–Liouville fractional integral of matrices. For some applications, it is often not necessary to compute exact solutions, approximate solutions are sufficient because sometimes computational efforts rapidly increase with the size of matrix functions. Our method is extended to find the exact and numerical solutions of the general system matrix convolution differential equations, the way exists which transform the coupled matrix differential equations into forms for which solutions may be readily computed. Finally, several systems are solved by the new and other approaches and illustrative examples are also considered. As usual, the notations A{1}(t) and det A(t) are the inverse and determinant of matrix function A(t), respectively, with respect to convolution. The notations AT(t) and Vec A(t) are transpose and vector-operator of matrix function A(t), respectively. The term ‘‘Vec A(t)’’ will be used to transform a matrix A(t) into a vector by stacking its column one underneath.The notations d(t) and Qn(t) = Ind(t) are the Dirac delta function and Dirac identity matrix, respectively, where In is the identity scalar matrix of order n · n. Finally, the notations A(t) ~ B(t) and A(t) * B(t) are the convolution and Kronecker convolution products, respectively. Finally, A B stands for the Kronecker product ; A B = [aijB]ij. 2. Operational matrices for integration and differentiation 2.1. Taylor series operational matrix for integration A function f(t) that is analytic function at the neighborhood of the point t0 can be expanded in the following formula: 1 X an un ðtÞ; ð1Þ f ðtÞ ¼ n¼1
n and un(t) = tn. where an ¼ n!1 d dtf ð0Þ n To obtain an approximate expression of the analytic function f(t) we may truncate the series (1) up to the (r + 1)th term: r X f ðtÞ ’ an un ðtÞ: ð2Þ n¼1
By defining the coefficient vector cT as: cT ¼ ða0 ; . . . ; ar Þ and the power series basis vector uT(t) as: uT ðtÞ ¼ ðu0 ðtÞ; . . . ; ur ðtÞÞ:
ð3Þ ð4Þ
Eq. (2) can written in the following compact form: f ðtÞ ’ cT uðtÞ:
ð5Þ
The basis functions un(t) satisfy the following relation: un ðtÞ ¼ tun1 ðtÞ: Now, one can easily show that Z t tnþ1 t 1 un ðtÞ ¼ u ðtÞ: un ðxÞ dx ¼ n þ 1 nþ1 nþ1 nþ1 0
ð6Þ
ð7Þ
252
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
On the basis of Eq. (2), the definite integral of analytic function f(t) in the interval (0, t) may be approximated as: Z t r r X X 1 unþ1 ðtÞ ¼ f ðxÞ dx ’ an bn unþ1 ðtÞ; ð8Þ nþ1 0 n¼1 n¼1 . where b0 ¼ 0; b1 ¼ a0 ; b2 ¼ a21 ; . . . ; br ¼ ar1 r In this equation the term br+1ur+1 has been truncated. Eq. (8) can also written as: Z t f ðxÞ dx ’ cT P uðtÞ;
ð9Þ
0
where P is called the Taylor 2 0 1 0 0 0 6 6 0 0 12 0 0 6 6 6 0 0 0 13 0 6 6 6 P ¼ 6 ... ... ... ... ... 6 6 60 0 0 0 0 6 6 60 0 0 0 0 4 0 0
0
0
series operational matrix for integration: 3 0 0 0 7 0 0 07 7 7 0 0 07 7 7 .. 7 .. .. .. : .7 . . . 7 7 1 0 r1 07 7 7 17 0 0 r 5
0
0
0
0
ð10Þ
ðrþ1Þðrþ1Þ
The above expression indicates that the Taylor series operational matrix for integration is uniquely simple. If we integrate the function f(t) twice in the interval (0, t), we have Z tZ t r r X X 1 1 2 f ðxÞðdxÞ ’ an cn unþ1 ðtÞ; ð11Þ unþ2 ðtÞ ¼ nþ1 nþ2 0 0 n¼1 n¼1 ar1 . where c0 ¼ c1 ¼ 0; c2 ¼ a0 ; c3 ¼ a21 ; . . . ; cr ¼ rðr1Þ Therefore, Eq. (11) can also written as: Z tZ t 2 f ðxÞðdxÞ ’ cT QuðtÞ 0
ð12Þ
0
and Q is called the operational matrix for double integration: 3 2 0 0 12 0 0 0 0 0 0 7 6 6 0 0 0 16 0 0 0 0 0 7 7 6 7 6 6 0 0 0 0 121 0 0 0 0 7 7 6 7 6 .. 7 .. .. .. 6 .. .. .. .. .. .. 7 6. . . . . . . . . . 7: 6 7 6 1 60 0 0 0 0 0 0 0 7 ðr1Þðr2Þ 7 6 7 6 60 0 0 0 0 0 0 1 7 0 6 rðr1Þ 7 7 6 60 0 0 0 0 0 0 0 0 7 5 4 0
0
0
0
0
0
0
0
ð13Þ
0
2
It easily to show that Q = P . Therefore, Eq. (12) may now written as: Z tZ t 2 f ðxÞðdxÞ ’ cT P 2 uðtÞ: 0
0
ð14Þ
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
253
Now we can extend Eq. (14) to the case of k multiple integrations as: Z
t
0
Z
t k
f ðxÞðdxÞ ’ aT P k uðtÞ:
ð15Þ
0
2.2. Generalized block pulse operational matrix for integration The block pulse functions form a complete set of orthogonal functions [10] which defined on the interval [0, b) by b 6 t < mi b; 1 i1 m wi ðtÞ ¼ ð16Þ 0 elsewhere . . . for i = 1, 2, . . . , m. It is also known that for any absolutely integrable function f(t) on [0, b) can be expanded in block pulse functions: f ðtÞ ’ nT /m ðtÞ;
ð17Þ
where the mean-square error of approximation is minimized, nT = [f1, f2, . . . , fm] and /m(t) = [w1(t), w2(t), . . . , wm(t)]. Here m fi ¼ b
Z
b 0
Z i m ðmÞb f ðtÞwi ðtÞ dt ¼ f ðtÞwi ðtÞ dt: b ði1 m Þb
ð18Þ
There are many different starting points for the discussion of classical fractional calculus [8]. One begins with a generalization of repeated integration. If f(t) is absolutely integrable on [0, b), it can be found that in [8,10] Z t Z tn Z t2 Z t Z t3 1 1 n1 tn1 f ðtÞ; dtn dtn1 dt2 f ðt1 Þdt1 ¼ ðt t1 Þ f ðt1 Þdt1 ¼ ð19Þ ðn þ 1Þ! ðn þ 1Þ! 0 0 0 0 0 where n = 1, 2, . . . , and 0 6 t < b. On writing C(n) = (n 1)!, an immediate generalization in the form of the operation Ia defined for a > 0 is Z t 1 1 a1 a1 a t f ðtÞ; 0 6 t < b; ðI f ÞðtÞ ¼ ðt t1 Þ f ðt1 Þdt1 ¼ ð20Þ CðaÞ 0 CðaÞ Rt a1 where C(a) is the Gamma function and ta1 f ðtÞ ¼ 0 f ðt t1 Þ ðt1 Þdt1 is called the convolution product of a1 t and f(t). Eq. (20) is called the Riemann–Liouville fractional integral of order a for the function f(t). Now if f(t) is expanded in block pulse functions, as shown in Eq. (17), the Riemann–Liouville fractional integral becomes ðI a f ÞðtÞ ¼
1 a1 1 t f ðtÞ ’ nT fta1 /m ðtÞg: CðaÞ CðaÞ
ð21Þ
Thus if ta1 * /m(t) can be integrated, then expanded in block pulse functions, the Riemann–Liouville fractional integral is solved via the block pulse functions. Let f1(t) = ta1, f2(t) = wi(t), a > 0, then 1 a1 1 t wi ðtÞ ¼ f1 ðtÞ f2 ðtÞ: CðaÞ CðaÞ Taking the Laplace transform of Eq. (22) gives 1 a1 1 1 t wi ðtÞ ¼ £ f1 ðtÞ f2 ðtÞ ¼ F 1 ðsÞF 2 ðsÞ; £ CðaÞ CðaÞ CðaÞ
ð22Þ
ð23Þ
254
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
where CðaÞ ; sa o 1 d 1 n i1 i wi ðtÞ , e½ m bs e½mbs : F 2 ðsÞ ¼ £fwi ðtÞg ¼ £ s dt s
F 1 ðsÞ ¼ £fta1 g ¼
Since C(a + 1) = aC(a), thus Eq. (23) can be written as: o 1 a1 1 Cða þ 1Þ n ½i1 i m bs e½mbs : t wi ðtÞ ¼ £ e CðaÞ Cða þ 1Þ saþ1 Taking the inverse Laplace transform of Eq. (25) yields a a Z t 1 1 i1 i1 i i t b u t b t b u t b ; ðt t1 Þa1 wi ðt1 Þdt1 ¼ CðaÞ 0 Cða þ 1Þ m m m m
ð24Þ
ð25Þ
ð26Þ
where u(t) is the unit step function. Further, it can be derived from Eq. (18) that ta uðtÞ ’ ½c1 ; c2 ; . . . ; cm /m ðtÞ ¼ C T /m ðtÞ;
ð27Þ
where ci ¼
a aþ1 b i ði 1Þaþ1 : m aþ1
ð28Þ
Moreover, from the disjoint property of the block pulse functions [5] wi ðtÞ i ¼ j; wi ðtÞwj ðtÞ ¼ 0 i 6¼ j: It is obvious that a i1 i1 b u t b ’ ½0; 0; . . . ; 0; c1 ; c2 ; . . . ; cmiþ1 /m ðtÞ; t m m a i i t b u t b ’ ½0; 0; . . . ; 0; c1 ; c2 ; . . . ; cm1 /m ðtÞ: m m Thus Eq. (26) becomes Z t 1 1 a1 ½0; 0; . . . ; 0; c1 ; c2 c1 ; . . . ; cmiþ1 cmi /m ðtÞ ðt t1 Þ wi ðt1 Þdt1 ’ CðaÞ 0 Cða þ 1Þ for i = 1, 2, . . . , m. Here, a b 1 aþ1 aþ1 ½raþ1 2ðr 1Þ þ ðr 2Þ cr cr1 ¼ m aþ1
a 1 for r = 2, 3, . . . , m i + 1 and c1 ¼ mb aþ1 . Now we can also write Eq. (32) as: a Z t 1 b 1 a1 ½0; 0; . . . ; 0; n1 ; n2 ; . . . ; nmiþ1 /m ðtÞ; ðt t1 Þ wi ðt1 Þdt1 ’ CðaÞ 0 m Cða þ 2Þ
ð29Þ
ð30Þ ð31Þ
ð32Þ
ð33Þ
ð34Þ
where n1 ¼ 1;
np ¼ paþ1 2ðp 1Þaþ1 þ ðp 2Þaþ1
ðp ¼ 2; 3; . . . ; m i þ 1Þ:
Finally, for i = 1, 2, . . . , m, Eq. (34) can be written as: Z t 1 a1 ðt t1 Þ /m ðt1 Þdt1 ’ F a /m ðtÞ; CðaÞ 0
ð35Þ
ð36Þ
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
255
where 2
1
6 60 a 6 b 1 60 Fa ¼ 6 m Cða þ 2Þ 6 6 40 0
n2
n3
1
n2
0
1
0 0
0 0
.. . 0
nm
3
7 nm1 7 7 nm2 7 7 .. 7 7 . 5 1
ð37Þ
is called the generalized block pulse operational matrix for integration. For example, if a = b = 1, we then have n1 = 1, np = p2 2 (2 1)2 + (p 2)2 = 2, for all p = 2, 3, . . . , m and 2 Z
t
0
3
1
2 2
2
60 6 1 6 60 /m ðt1 Þdt1 ’ F 1 /m ðtÞ ¼ 2m 6 6 40 0
1 2 0 1
27 7 7 2 7/m ðtÞ: 7 .. 7 .. . .5 0 1
0 0 0 0
The generalized block pulse operational matrices Da for differentiation can be derived simply by inverting the Fa matrices. Let 2
Da ¼ F 1 a
1 6 60 6 ma 6 ¼ Cða þ 2Þ6 0 6 b 6 40 0
n2 1
n3 n2
0
1
0 0
0 0
.. . 0
31 2 nm d1 7 6 nm1 7 60 7 6 ma 6 nm2 7 Cða þ 2Þ6 0 7 ¼ 7 6 b .. 7 6 . 5 40 0 1
d2
d1
d1 0
d2 d1
0
0
0
0
dm
3
7 d m1 7 7 d m2 7 7: .. 7 .. 7 . 5 . 0 d1 ð38Þ
Thus 2
1 6 60 6 60 6 6 6 40 0
n2 1
n3 n2
0
1
0 0
0 0
.. . 0
32 nm d1 76 nm1 7 6 0 76 6 nm2 7 7:6 0 7 .. 7 6 6 . 54 0 1
0
d2
d3
d1 0
d2 d1
0
0
0
0
dm
3
7 d m1 7 7 d m2 7 7 ¼ I m; .. 7 .. 7 . 5 . 0 d1
where Im is the identity matrix of order m · m. After equating both sides of Eq. (39), we get m X ni d miþ1 : d 1 ¼ 1; d 2 ¼ n2 d 1 ; . . . ; d m ¼
ð39Þ
ð40Þ
i¼2
Note that if a = 0, then F0 = D0 = Im which corresponds to s0 = 1 in the Laplace domain. It is obvious that the Fa and Da matrices perform as sa and sa in the Laplace domain and as fractional (operational) integrators and differentiators in the time domain. For example, let a = 0.5, m = 4, b = 1, the operational matrices F0.5 and D0.5 which correspond to s0.5 and 0.5 s are computed p below: ffiffiffi pffiffiffi pffiffiffi pffiffiffi pffiffiffi n1 = 1, n2 ¼ 2 2 2 ¼ 0:8284, n3 ¼ 3 3 4 2 þ 1 ¼ 0:5393, p n4ffiffiffi¼ 8 6 3 þ 2 2 ¼ 0:4631, d 1 ¼ 1, d2 = 0.8284, d3 = 0.1470, n4 = 0.1111, Cð2:5Þ ¼ 0:75Cð0:5Þ ¼ 0:75 p ¼ 1:3293 and
256
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
2
0:3761 0:3116
6 6 F 0:5 ¼ 6 6 4 2 D0:5
0:2028
0
0:3761
0:3116
0
0
0:3761
0
0
0
0:1640
7 0:2028 7 7 0:3116 7 5 0:3761
2:6857
2:2025
0:2028
0
2:6857
2:2025
0
0
2:6857
0
0
0
6 6 ¼6 6 4
3
0:2954
3
7 0:3098 7 7: 2:2025 7 5 2:6857
3. Convolution product and Riemann–Liouville fractional integral of matrices Definition 1. Let A(t) = [fij(t)] and B(t) = [gij(t)] be m · m absolutely integrable matrices on [0, b). The convolution and Kronecker convolution product s are matrix functions defined by (i) Convolution product: AðtÞ BðtÞ ¼ ½hij ðtÞ; hij ðtÞ ¼
m Z X k¼1
t
fik ðt t1 Þgkj ðt1 Þdt1 ¼
0
m X
fik ðtÞ gkj ðtÞ:
ð41Þ
k¼1
(ii) Kronecker convolution product AðtÞ ~ BðtÞ ¼ ½fij ðtÞ BðtÞij :
ð42Þ
Here, fij(t) * B(t) is the ijth submatrix of order m · m and A(t) ~ B(t) is of order m2 · m2. h i Definition 2. Let AðtÞ ¼ Cða1ij Þ taij 1 and B(t) = [gij(t)] be m · m absolutely integrable matrices on [0, b). The Riemann–Liouville Fractional Integral of A(t) and B(t) is a matrix functions defined by ðI aij BÞðtÞ ¼ AðtÞ BðtÞ;
ð43Þ
aij > 0:
Definition 3. Let A(t) = [fij(t)] be m · m absolutely integrable matrices on [0, b). The determinant, inverse and k-power of A(t) respect to the convolution are defined by (i) Determinant: m X ð1Þjþ1 f1j ðtÞ D1j ; det AðtÞ ¼
ð44Þ
j¼1
where Dij the determinant of the (m 1) · (m 1) matrix function obtained from A(t) by deleting row i and column j of A(t). We call Dij(t) the minor A(t) of corresponding to the entry of fij(t) of A(t).For example, if A(t) = [fij(t)] be 2 · 2 absolutely integrable matrices on [0, b), then det AðtÞ ¼ f11 ðtÞ f22 ðtÞ f12 ðtÞ f21 ðtÞ: (ii) Inversion: Af1g ðtÞ ¼ ½qij ðtÞ;
qij ðtÞ ¼ ðdet AðtÞÞ
{1}
f1g
{1}
adj AðtÞ:
ð45Þ {1}
{1}
Here, (det A(t)) exists, (det A(t)) (t) * A(t) = A(t) * A * det A(t) = d(t) and note that A Qm(t), where Qm(t) = Imd(t) is the Dirac identity matrix and d(t) is the Dirac identity function. (iii) k- power convolution product: m X fmg fmg fm1g fir ðtÞ frj ðtÞ: Afkg ðtÞ ¼ ½fij ðtÞ; fij ðtÞ ¼ r¼1
(t) =
ð46Þ
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
257
Theorem 4. Let A(t), B(t) and C(t) be n · n absolutely integrable matrices on [0, b).Then for any scalars a and b ðiÞ ½aAðtÞ þ bBðtÞ CðtÞ ¼ aAðtÞ CðtÞ þ bBðtÞ CðtÞ;
ð47Þ
ðiiÞ ðAðtÞ BðtÞÞ CðtÞ ¼ AðtÞ ðBðtÞ CðtÞÞ; ðiiiÞ AðtÞ Qn ðtÞ ¼ Qn ðtÞ Qn ðtÞ AðtÞ ¼ AðtÞ;
ð48Þ ð49Þ
T
Qn ðtÞ ¼ I n dðtÞ;
T
T
ðivÞ ½AðtÞ BðtÞ ¼ B ðtÞ A ðtÞ; ðvÞ ½AðtÞ BðtÞ
f1g
¼B
f1g
ðtÞ A
f1g
ð50Þ ðtÞ; if A
f1g
ðtÞ and B
f1g
ðtÞ are exist:
ð51Þ
Proof. Straightforward by the definition of the convolution product of matrices.
h
Lemma 5. Let A(t) = [fij(t)] and B(t) = [gij(t)] be m · m Semi-Markov Kernel matrices on [0, b). Then for fixed t ðAðtÞ BðtÞÞði; jÞ 6 ðAðtÞBðtÞÞði; jÞ:
ð52Þ
Proof. Since t 2 [0, b) (fixed), then ðAðtÞ BðtÞÞði; jÞ ¼
m X
fik ðtÞ gkj ðtÞ 6
k¼1
m X
fik ðtÞgkj ðtÞ ¼ ðAðtÞBðtÞÞði; jÞ:
k¼1
Theorem 6. Let A(t) = [fij(t)] and B(t) = [gij(t)] be m · m Semi-Markov Kernel matrices on [0, b). Then for any matrix norm kÆk and positive integer k kAðtÞ BðtÞk 6 kAðtÞkkBðtÞk; kAðtÞ
fkg
ð53Þ
k
k 6 kAðtÞk :
ð54Þ
Proof. Since kÆk is matrix norm, then by Eq. (52), we have kAðtÞ BðtÞk 6 kAðtÞBðtÞk 6 kAðtÞkkBðtÞk: The proof of other inequality is by induction on k and Eq. (53), we have k
kAfkg ðtÞk ¼ kAfk1g ðtÞ AðtÞk 6 kAfk1g ðtÞkkAðtÞk 6 6 kAðtÞk :
Theorem 7. Let A(t) = [fij(t)], B(t) = [gij(t)] be m · m Laplace transformable matrices. Then ðiÞ £fAðtÞ BðtÞg ¼ £fAðtÞg£fBðtÞg;
ð55Þ
ðiiÞ £fAðtÞ ~ BðtÞg ¼ £fAðtÞg £fBðtÞg:
ð56Þ
Proof. (i) Straightforward by the convolution Theorem. (ii) Suppose that £{fij(t)} = Fij(s) and £ {B(t)} = G(s), then by part (i), we have £fAðtÞ ~ BðtÞg ¼ £½fij ðtÞ BðtÞij ¼ ½F ij ðsÞGðsÞ ¼ £fAðtÞg £fBðtÞg:
Theorem 8. Let A(t), B(t) and C(t) be n · n absolutely integrable matrices on [0, b). Then ½AðtÞ ~ BðtÞ ½CðtÞ ~ DðtÞ ¼ ½AðtÞ CðtÞ ~ ½BðtÞ DðtÞ:
ð57Þ
258
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
Proof. The (i, j)th block of [A(t) ~ B(t)]*[C(t) ~ D(t)] is obtained by taking the convolution product of ith row block of A(t) ~ B(t) and the jth column block of C(t) ~ D(t),i.e., 2 6 6 ½fi1 ðtÞ BðtÞ fin ðtÞ BðtÞ 6 4
h1j ðtÞ DðtÞ .. .
3 n 7 X 7 ðfik ðtÞ hkj ðtÞ BðtÞ DðtÞÞ 7¼ 5 k¼1
hnj ðtÞ DðtÞ and the (i, j)th block of the right hand side [A(t) * C(t)] ~ [B(t) * D(t)] obtained by the definition of the Kronecker convolution product xij(t)*(B(t) * D(t)) where xij is the (i, j)th element in A(t) * C(t). But the rule of convolution product xij ðtÞ ¼
n X
ðfik ðtÞ hkj ðtÞÞ:
k¼1
Lemma 9. Let u(t) and v(t) be m-vectors. Then VecðuðtÞ vT ðtÞÞ ¼ vðtÞ ~ uðtÞ:
ð58Þ
Proof. Straightforward by the definitions of convolution and Kronecker convolution products. h Theorem 10. Let A(t), X(t), and B(t) be m · m absolutely integrable matrices on [0, b).Then ðiÞ Vec½AðtÞ X ðtÞ BðtÞ ¼ ½BT ðtÞ ~ AðtÞ Vec X ðtÞ;
ð59Þ
T
ðiiÞ Vec½AðtÞX ðtÞ BðtÞ ¼ ½B ðtÞ AðtÞ Vec X ðtÞ:
ð60Þ
Proof. Let X(t) be contains x1(t), . . . , xm(t) vectors each is of order m ·P 1 and let ej(t) be m · 1a vector of zeros except for a d(t) in the jth position, for each j = 1, 2 . . . , m, then X ðtÞ ¼ mj¼1 xj ðtÞ eTj ðtÞ. Now, since A(t) * xj(t) and BT(t) * ej(t) are vectors of order m · 1, then we have ( VecfAðtÞ X ðtÞ BðtÞg ¼ Vec
m X
) AðtÞ xj ðtÞ
eTj ðtÞ
j¼1
¼
m X
BðtÞ
¼
m X
T
VecfðAðtÞ xj ðtÞÞ ðBT ðtÞ ej ðtÞÞ g
j¼1
m X f½BT ðtÞ ej ðtÞ ~ ðAðtÞ xj ðtÞÞg ¼ f½BT ðtÞ ~ AðtÞ ðej ðtÞ ~ xj ðtÞÞg
j¼1
j¼1
¼ ½BT ðtÞ ~ AðtÞ
m X
Vecðxj ðtÞ ~ eTj ðtÞÞ ¼ ðBT ðtÞ ~ AðtÞÞ Vec
j¼1
m X
ðxj ðtÞ ~ eTj ðtÞÞ
j¼1
¼ ðBT ðtÞ ~ AðtÞÞ Vec X ðtÞ: Similarly, we can proof (ii).
h
4. Some applications 4.1. Solution of the state–space equations and optimal control problem using Taylor series and Kronecker product The state–space equation is given by x0 ðtÞ ¼ AxðtÞ þ BuðtÞ;
xð0Þ ¼ x0 ;
0 6 t 6 b;
ð61Þ
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
259
where xðtÞ 2 Rn , uðtÞ 2 Rm analytic and A, B are known constant matrices. The input vector u(t) may be expanded in Taylor series as follows: 3 2 3 3 2 2 h10 h11 h12 h1;r1 u0 ðtÞ u1 ðtÞ 7 6 7 7 6 6 6 u2 ðtÞ 7 6 h20 h21 h22 h2;r1 7 6 u1 ðtÞ 7 7 6 7 7 6 6 ð62Þ 6 uðtÞ ¼ 6 . 7 ¼ 6 . 7 ¼ H uðtÞ; .. 7 .. .. .. 7 6 .. 7 6 . 7 6 . . 5 4 . 5 . . . 4 . 5 4 . um ðtÞ
hm0
hm1
hm;r1
hm2
ur1 ðtÞ
where H is a known constant matrix. Similarly, the state vector x(t) may also be expanded in Taylor series as follows: 3 2 3 32 2 f10 f11 f12 f1;r1 x1 ðtÞ u0 ðtÞ 7 6 7 76 6 6 x2 ðtÞ 7 6 f20 f21 f22 f2;r1 7 6 u1 ðtÞ 7 7 6 7 76 6 xðtÞ ¼ 6 . 7 ¼ 6 . :6 7 ¼ ½f0 ; f1 ; f2 ; . . . ; fr1 :uðtÞ ¼ F uðtÞ; .. 7 .. .. .. 7 6 .. 7 6 . 7 6 . . 54 . 5 . . . 4 . 5 4 . xn ðtÞ
fn0
fn1
fn2
fn;r1
ur1 ðtÞ
where F contains the unknown coefficients of the Taylor expansion of the state vector x(t). Next by integrating the space–state equation in Eq. (61), we have Z t Z t xðtÞ xð0Þ ¼ A xðrÞ dr þ B uðrÞ dr: 0
ð63Þ
ð64Þ
0
Introducing Eqs. (62), (63) in (64) and using the integration property of the operational matrix P expressed by (15), we obtain ðF AFP ÞuðtÞ ¼ ðBHP þ EÞuðtÞ;
ð65Þ
where E = [x0 0 0 . . . 0]. In Eq. (65), the matrix BHP + E is known; therefore, we have the following system for the unknown matrix F: AFP D;
D ¼ BHP E:
ð66Þ
This system may be written in a simpler form by using the Kronecker product, actually we have Mf ¼ ðA P T I nr Þf ¼ d;
ð67Þ
T
where M = P A Inr is a nr · nr matrix and 2 3 2 f0 d0 6 f 7 6 d 6 1 7 6 1 7 6 f ¼ Vec F ¼ 6 6 .. 7; d ¼ Vec D ¼ 6 .. 4 . 5 4 . fr1 i.e., fi and di are the ith columns of F and 2 pA pA 6 pA pA 6 M ¼ P T A I nr ¼ 6 .. .. 6 .. 4 . . . pA pA
3 7 7 7; 7 5
d r1 D, respectively. Here, 3 pA pA 7 7 .. 7 7 I nr . 5 pA
is called the Kronecker operational matrix, where Inr is an identity matrix of order nr · nr. The solution of Eq. (67) is f = M1d = (PT A Inr)1d.
260
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
The main advantage of this approach is that matrix M = PT A Inr due to the form of the operational matrix P, can be easily proven lower triangular. Therefore, one does not have to inverse the matrix M to find the solution. Example 11. Consider the following state–space equation 0 1 0 xðtÞ ¼ xðtÞ þ uðtÞ; 2 3 1 0 , u(t) = 2t + 3, 0 6 t 6 b. where xð0Þ ¼ 1 If we choose r = 4 terms, then h = [3, 2, 0, 0]T, u(t) = [u0(t), u1(t), u2(t), u3(t)] = [1, t, t2, t3]. Now 1 0 0 0 D ¼ BHP E ¼ 1 3 1 0 T d ¼ Vec D ¼ ½1; 1; 0; 3; 0; 1; 0; 0 2 1 0 0 0 0 0 0 6 0 1 0 0 0 0 0 2 3 6 6 0 0 0 I 2 1 1 0 0 0 0 6 0 6 6 A I 7 6 0 0 2 3 0 1 0 0 0 2 6 7 6 M ¼ P T A I8 ¼ 6 7¼6 A I 0 4 0 5 6 0 0 0 1=2 1 0 0 2 2 6 A I 0 0 0 0 1 3=2 0 1 0 6 2 3 6 4 0 0 0 0 0 1=3 1 0 0 0 0 2=3 1 0 Now, f = M1d = [0, 1, 1, 0, 0, 0, 0, 0]T. Therefore, x1 ðtÞ 0 1 0 xðtÞ ¼ ¼ ½f0 ; f1 ; f2 ; f3 :uðtÞ ¼ 1 0 0 x2 ðtÞ
0 0
uðtÞ ¼
t 1
3 0 0 7 7 7 0 7 7 0 7 7: 0 7 7 7 0 7 7 0 5 1
:
The optimal control problem for the system in Eq. (61) with the performance index Z 1 b T ½x Qx þ uT Rudr Y ¼ 2 0
ð68Þ
has the well known solution u ¼ R1 BT kðtÞ; where k(t) satisfies the following canonical equation: # 0 " xðtÞ x ðtÞ A BR1 BT ; ¼ kðtÞ k 0 ðtÞ Q AT with the boundary conditions x(0) = x0, and k(b) = 0. To simply the solution of the system (70), we change the independent variable t as s = b t. We now have # 0 " xðsÞ x ðsÞ A BR1 BT xðsÞ ¼ N ; ¼ kðsÞ kðsÞ k 0 ðsÞ Q AT with the boundary conditions x(t = 0) = x(s = b) = x0 and k(t = b) = k(s = 0) = 0. Integration of Eq. (71) from 0 to s yields: " Rs # xðrÞdr xðsÞ xð0Þ 0 : ¼ N R s kðrÞdr kðsÞ kð0Þ 0
ð69Þ
ð70Þ
ð71Þ
ð72Þ
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
261
If we assume that the functions x(s) and k(s) are analytic, then we may expand them in Taylor series as follows: xðsÞ ¼ W uðsÞ: ð73Þ kðsÞ The matrix W is the 2n · r Taylor coefficient matrix that must be determined. With the use of Eqs. (73), (72) can be written as: W uðsÞ þ NWP uðsÞ ¼ SuðsÞ; ð74Þ where P is the operational matrix expressed by (15). By equating the corresponding terms of the Taylor series, we obtain the following matrix equation: W þ NWP ¼ S: ð75Þ Applying the Vec-notation and Kronecker product, we have Vec W þ VecðNWP Þ ¼ Vec W þ ðP T N ÞVec W ¼ Vec S:
ð76Þ
Now, Eq. (75) may be written as: Tw ¼ s
where T ¼ I þ P T N ;
ð77Þ
where T is a 2nr · 2nr square matrix. Also, xðs ¼ 0Þ 0 0 0 S¼ 0 0 0 0 and
2 6 6 w¼6 4
w0 w1 .. .
3
2
7 7 7; 5
6 6 s¼6 4
s0 s1 .. .
3 7 7 7 5
wr1 sr1 are the vectors formed by the vector elements wi, si that are the columns of matrices W and S, respectively. Due to the structure of operational matrix of integration P, the matrix T formed by the kronecker product of PT and N, may be easily proved lower triangular. This fact simplifies immensely the solution of the system in Eq. (77) that generally has 2nr equations. 4.2. Solution of the general coupled convolution matrix differential equations using Kronecker Products The General Coupled Convolution Matrix Differential Equations is given by X 01 ðtÞ ¼ A11 ðtÞ X 1 ðtÞ B11 ðtÞ þ A12 ðtÞ X 2 ðtÞ B12 ðtÞ þ þ A1p ðtÞ X p ðtÞ B1p ðtÞ X 02 ðtÞ ¼ A21 ðtÞ X 1 ðtÞ B21 ðtÞ þ A22 ðtÞ X 2 ðtÞ B22 ðtÞ þ þ A2p ðtÞ X p ðtÞ B2p ðtÞ ; .. . X 0p ðtÞ ¼ Ap1 ðtÞ X 1 ðtÞ Bp1 ðtÞ þ Ap2 ðtÞ X 2 ðtÞ Bp2 ðtÞ þ þ App ðtÞ X p ðtÞ Bpp ðtÞ
ð78Þ
where Aij(t), Bij(t) are given m · m matrices with entries in are a restricted set of functions on [0, b), which is closed under convolution and isomorphic via the Laplace transforms to the proper rational functions with natural multiplication, Xi(t) unknown matrix functions to be solved and Xi(0) = Ci (i, j = 1, 2, . . . , p). If we use the Vec- notation and the Kronecker convolution product as in Eq. (59) to the system in Eq. (78), we have 3 2 3 2 T B11 ðtÞ ~ A11 ðtÞ BT12 ðtÞ ~ A12 ðtÞ BT1p ðtÞ ~ A1p ðtÞ 2 Vec X 1 ðtÞ 3 Vec X 01 ðtÞ 6 Vec X 0 ðtÞ 7 6 BT ðtÞ ~ A ðtÞ BT ðtÞ ~ A ðtÞ BT ðtÞ ~ A ðtÞ 76 Vec X ðtÞ 7 21 22 2p 2 76 6 7 6 21 2 22 2p 7 76 6 7¼6 ð79Þ 7: .. .. .. .. .. .. 74 6 7 6 5 . 5 4 5 4 . . . . . Vec X 0p ðtÞ Vec X p ðtÞ BTp1 ðtÞ ~ Ap1 ðtÞ BTp2 ðtÞ ~ Ap2 ðtÞ BTpp ðtÞ ~ App ðtÞ
262
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
This system may be written in a simpler form by x0 ðtÞ ¼ H ðtÞ xðtÞ;
xðtÞ ¼ c;
ð80Þ
where 2
3
Vec X 01 ðtÞ
2
Vec X 1 ðtÞ
3
7 7 6 6 6 Vec X 2 ðtÞ 7 6 Vec X 02 ðtÞ 7 7 7 6 6 7 7; xðtÞ ¼ 6 x0 ðtÞ ¼ 6 7 7 6 6 .. .. 7 7 6 6 . . 5 5 4 4 0 Vec X p ðtÞ Vec X p ðtÞ 3 2 T B11 ðtÞ ~ A11 ðtÞ BT12 ðtÞ ~ A12 ðtÞ BT1p ðtÞ ~ A1p ðtÞ 7 6 6 BT ðtÞ ~ A ðtÞ BT ðtÞ ~ A ðtÞ BT ðtÞ ~ A ðtÞ 7 7 6 21 21 22 2p 22 2p 7 6 H ðtÞ ¼ 6 7: .. .. .. .. 7 6 7 6 . . . . 5 4 BTp1 ðtÞ ~ Ap1 ðtÞ
BTp2 ðtÞ ~ Ap2 ðtÞ
ð81Þ
ð82Þ
BTpp ðtÞ ~ App ðtÞ
Now if we use Laplace transform to Eq. (80), we have £fx0 ðtÞg ¼ £fH ðtÞ xðtÞg:
ð83Þ
Let Y(s) = £{x(t)}, G(s) = £{H(t)}, we have Y ðsÞ ¼ fsI GðsÞg
f1g
c;
sI > GðsÞ:
Taking the inverse Laplace transform of Eq. (84) yields ( ( f1g ) f1g !) GðsÞ 1 GðsÞ I xðtÞ ¼ £1 s I c ¼ £1 c ¼ I ðI I H ðtÞÞf1g c; s s s
ð84Þ
ð85Þ
where I is an identity scalar matrix. Suppose that Q(t) = I * H(t), then (I I * H(t)){1} = (I Q(t)){1} can be obtained by two ways, either by truncated series development or by explicit inversion within the above described convolution algebra. In the first case, we have n X f1g ¼Iþ Qfkg ðtÞ þ Rn ðtÞ: ð86Þ ðI QðtÞÞ k¼1
Here, fnþ1g
Qij
ðtÞ ¼
XZ k2N
t
fng
Qik ðuÞQkj ðt uÞ du
ð87Þ
0
and Q{1}(t) = Q(t). The total error is given by 1 X Qfkg ðtÞ ¼ Qfnþ1g ðtÞ ðI þ QðtÞ þ Qf2g ðtÞ þ Þ ¼ Qfnþ1g ðtÞ ðI QðtÞÞf1g : Rn ðtÞ ¼
ð88Þ
k¼nþ1
Now for any matrix norm satisfying: n
kQfng ðtÞk 6 kQðtÞk ;
ð89Þ
then the total error due to the above truncation is also bounded by (89), if kQ(t)k < 1, then kRn ðtÞk 6
kQðtÞknþ1 : 1 kQðtÞk
ð90Þ
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
263
Thus for given maximum error, say e we have nþ16
lnfeð1 kQðtÞkÞg : ln kQðtÞk
ð91Þ
In the direct inversion method, we calculate (I Q(t)){1} directly: ðI QðtÞÞ
f1g
¼ ðdetðI QðtÞÞÞ
f1g
adjðI QðtÞÞ:
ð92Þ
Example 12. The following coupled matrix convolution equations: AðtÞ X ðtÞ þ Y ðtÞ BðtÞ ¼ CðtÞ; DðtÞ X ðtÞ þ Y ðtÞ EðtÞ ¼ F ðtÞ has a unique solution if and only if the matrix " # Qn ðtÞ ~ AðtÞ BT ðtÞ ~ Qm ðtÞ 2 M I2mn H¼ Qn ðtÞ ~ DðtÞ ET ðtÞ ~ Qm ðtÞ is non-singular; in this case the unique solution is given by Vec CðtÞ Vec X ðtÞ f1g ðtÞ ¼H Vec F ðtÞ Vec Y ðtÞ and the corresponding homogeneous matrix convolution equations AðtÞ X ðtÞ þ Y ðtÞ BðtÞ ¼ 0; DðtÞ X ðtÞ þ Y ðtÞ EðtÞ ¼ 0 has a unique solution: X(t) = Y(t) = 0. Because if we use the Vec-notation and Kronecker product, we have " # Vec X ðtÞ Vec CðtÞ Qn ðtÞ ~ AðtÞ BT ðtÞ ~ Qm ðtÞ ¼ : Vec Y ðtÞ Vec F ðtÞ Qn ðtÞ ~ DðtÞ ET ðtÞ ~ Qm ðtÞ This system has a unique solution if and only if the matrix H is non-singular. 4.3. Invert the rational and irrational transfer functions We will use generalized block pulse operational matrices Fa and Da that defined in Eqs. (37) and (38) to get the time functions of rational and irrational transfer functions. The irrational transfer functions are derived from distributed systems. It can be found in [3]. Consider any transfer function G(s) whose time function is denoted by g(t). Let 1 ð93Þ GðsÞ ¼ ZðsÞ; s where 1s is the Laplace transform of unit step function. The meaning of (93) is that g(t) can be derived by applying a unit step function to a new system whose transfer function is Z(s) = sG(s). Given the rational transfer function G(s), b1 sn1 þ b2 sn2 þ þ bn1 s þ bn GðsÞ ¼ : ð94Þ sn þ a1 sn1 þ þ an1 s þ an Then b1 sn þ b2 sn1 þ þ bn1 s2 þ bn s ¼ g1 þ g2 s1 þ g3 s2 þ þ gkþ1 sk þ ; ð95Þ ZðsÞ ¼ n s þ a1 sn1 þ þ an1 s þ an where gi can be derived by cross multiplications of (95) to yield the following linear recursive equation: g 1 ¼ b1 ;
g 2 ¼ b 2 a1 g 1 ;
g n ¼ bn
n1 X
ani gi ;
i¼1
gnþ1 ¼
n X i¼1
anþ1i gi ;
gnþr ¼
nþr1 X i¼1
anþri gi :
ð96Þ
264
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
Eq. (95) is the Maclaurin series of G(s). For irrational transfer functions G(s), Z(s) can always be expanded in Maclaurin series. This can be seen from the following examples. After associating each sa with Fa(a < 0) or Da(a > 0) operational matrices, the approximation of g(t) via the block pulse functions can be immediately derived and found in [3,10]. Example 13. Given the rational transfer function G(s), s4 þ 3s3 þ 2s2 þ 5s þ 1 1 ¼ ð1 2s1 þ 9s2 41s3 þ 192s4 Þ; þ 5s4 þ 3s3 þ 7s2 þ s þ 1 s ) gðtÞ ’ ½1; 1; . . . ; 1ðI 2F 1 þ 9F 2 41F 3 þ 192F 4 Þum ðtÞ: GðsÞ ¼
s5
For b = 5, m = 10, gðtÞ ’ ½0:7325; 0:5878; 0:6187; 0:6847; 0:7374; 0:7396; 0:6679; 0:5185; 0:3081; 0:0681u10 ðtÞ: Example 14. The Laplace transform of the Bessel function of the first kind and zero order J0(t) is 1 1 1 1 1 3 5 35 8 1 s2 þ s4 s6 þ s ; GðsÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 2 8 16 128 s2 þ 1 s 1 þ s2 s 1 3 5 35 F 4 um ðtÞ: ) gðtÞ ’ ½1; 1; . . . ; 1 I F 2 þ F 4 F 6 þ 2 8 16 128 For b = 8, m = 8, gðtÞ ’ ½0:9197; 0:5060; 0:0382; 0:3628; 0:3094; 0:0090; 0:2484; 0:2561u8 ðtÞ: Example 15. The Laplace transform of the generalized exponential function is p1ffi 1 1 s GðsÞ ¼ pffiffiffiffiffiffiffiffiffiffiffi ¼ ð1 þ s0:5 Þ1 ¼ ð1 s0:5 þ s1 s1:5 þ s2 s2:5 þ Þ; s sþ1 s ) gðtÞ ’ ½1; 1; . . . ; 1ðI F 0:5 þ F 1 F 1:5 þ F 2 F 2:5 þ Þum ðtÞ:
For b = 2.4, m = 4, gðtÞ ’ ½0:6201; 0:4448; 0:3742; 0:3306u4 ðtÞ: Example 16. The Laplace transform of the following irrational transfer function is 1 pffis 1 pffiffi pffis 1 0:5 1 1:5 1 2 1 2:5 1 3 se s sþ s s þ s s þ ; GðsÞ ¼ pffiffi e ¼ ¼ s s 2 6 24 120 s 1 1 1 1 D3 þ um ðtÞ: ) gðtÞ ’ ½1; 1; . . . ; 1 D0:5 D1 þ D1:5 D2 þ D2:5 2 6 24 120 For b = 5, m = 5, gðtÞ ’ ½0:3782; 0:3425; 0:1824; 0:0840; 0:1413u5 ðtÞ: Acknowledgements The present research has been partially supported by University Putra Malaysia (UPM) under the grant IRPA09-02-04-0259-EA001. References [1] Al Zhour Zeyad, Kilicman Adem, Some applications of the convolution and Kronecker products of matrices, in: Proceeding of the International Conference on Math. 2005, UUM, Kedah, Malaysia, 2005, pp. 551–562.
A. Kilicman, Z.A.A. Al Zhour / Applied Mathematics and Computation 187 (2007) 250–265
265
[2] T. Chen, B.A. Francis, Optimal Sampled-Data Control Systems, Springer, London, 1995. [3] C.F. Chen, Y.T. Tsay, T.T. Wu, Walsh operational matrices for fractional calculus and their applications to distributed systems, J. Frankin Inst. 303 (3) (1977) 267–284. [4] A.E. Gilmour, Circulant Matrix Methods for the Numerical solutions of Partial Differential Equations by FFT Convolutions, New Zealand, 1987. [5] K. Maleknejad, M. Shaherezaee, H. Khatami, Numerical solution of integral equations systems of second kind by Block Pulse Functions, Appl. Math. Comput. 166 (2005) 15–24. [6] S.G. Mouroutsos, P.D. Sparis, Taylor Series approach to system identification, analysis and optimal control, J. Frankin Inst. 319 (3) (1985) 359–371. [7] L. Nikolaos, Dependability analysis of Semi-Markov system, Reliab. Eng. Syst. Safety 55 (1997) 203–207. [8] B. Ross, Fractional Calculus and its Applications, Springer-Verlag, Berlin, 1975. [9] H. Sumita, The Matrix Laguerre Transform, Appl. Math. Comput. 15 (1984) 1–28. [10] C.-H. Wang, On the generalization of Block Pulse Operational matrices for fractional and operational calculus, J. Frankin Inst. 315 (2) (1983) 91–102.