<<

Computers and Mathematics with Applications 56 (2008) 2941–2947

Contents lists available at ScienceDirect

Computers and Mathematics with Applications

journal homepage: www.elsevier.com/locate/camwa

A generalization of Fourier trigonometric

Mohammad Masjed-Jamei a,b,∗, Mehdi Dehghan a a Department of Applied Mathematics, Faculty of Mathematics and Computer Science, Amirkabir University of Technology, No 429, Hafez Ave, Tehran, Iran b Department of Applied Mathematics, K.N.Toosi University of Technology, P.O. Box 16315-1618, Tehran, Iran article info a b s t r a c t

Article history: In this paper, by using the extended Sturm–Liouville theorem for symmetric functions, we Received 30 August 2007 introduce the differential equation Received in revised form 31 May 2008  +  Accepted 10 July 2008 00 n 2 a(a 1) n Φ (t) + n + a(1 − (−1) )/2 − (1 − (−1) )/2 Φn(t) = 0, n cos2 t Keywords: { }∞ Extended Sturm–Liouville theorem for as a generalization of the differential equation of trigonometric sin nt n=1 and { }∞ = symmetric functions cos nt n=0 for a 0 and obtain its explicit solution in a simple trigonometric form. Symmetric orthogonal functions We then prove that the obtained of solutions is orthogonal with respect to the Norm square value constant weight function on [0, π] and compute its norm square value explicitly. One of the Fourier trigonometric sequences important advantages of this generalization is to find some new infinite series. A practical Hypergeometric functions example is given in this sense. © 2008 Elsevier Ltd. All rights reserved.

1. Introduction

Many sequences of special functions are solutions of a usual Sturm–Liouville problem [1,2]. For instance, the well-known { }∞ { }∞ trigonometric sequences sin nx n=1 and cos nx n=0, which respectively generate the Fourier sine and cosine series, are solutions of a usual Sturm–Liouville equation in the form 00 + 2 = ∈ [ ] Φn (x) n Φn(x) 0, x 0, π . (1) It is clear that the interval [0, π] in equation (1) can be transformed to any other arbitrary interval, say [−l, l] with period 2l, by a simple linear transformation. On the other hand, most special functions applied in physics, mathematics and engineering, satisfy a symmetry relation as n + Φn(−x) = (−1) Φn(x) ∀n ∈ Z . (2) Recently in [3], we have presented a key theorem by which one could generalize the usual Sturm–Liouville problems with symmetric solutions. We have shown that the solutions corresponding to an extended Sturm–Liouville equation are orthogonal [4,5] with respect to an even weight function on a symmetric interval. In other words:

n Theorem 1.1 ([3]). Let Φn(x) = (−1) Φn(−x) be a sequence of symmetric functions that satisfies a differential equation of the form 00 + 0 + + + − − n  = A(x)Φn (x) B(x)Φn(x) λnC(x) D(x) (1 ( 1) )E(x)/2 Φn(x) 0, (3)

∗ Corresponding author. E-mail addresses: [email protected], [email protected] (M. Masjed-Jamei), [email protected] (M. Dehghan).

0898-1221/$ – see front matter © 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.camwa.2008.07.023 2942 M. Masjed-Jamei, M. Dehghan / Computers and Mathematics with Applications 56 (2008) 2941–2947 where A(x), B(x), C(x), D(x) and E(x) are real functions and {λn} is a sequence of constants. If A(x), C(x), D(x) and E(x) are even functions, C(x) is positive and B(x) is an odd function then Z α Z α   ∗ ∗ 0 if n 6= m, W x x x dx = W x 2 x dx with = (4) ( )Φn( )Φm( ) ( )Φn ( ) δn,m δn,m = −α −α 1 if n m, where Z 0  Z  ∗ B(x) − A (x) C(x) B(x) W (x) = C(x) exp dx = exp dx . (4.1) A(x) A(x) A(x) Of course, the weight function W ∗(x) must be positive and even on [−α, α] and x = α must be a root of the function

Z B(x) − A0(x)  Z B(x)  A(x)K(x) = A(x) exp dx = exp dx , (4.2) A(x) A(x) i.e. A(α)K(α) = 0. In this sense, since K(x) = W ∗(x)/C(x) is an even function then A(−α)K(−α) = 0 automatically. By using this theorem fortunately many new symmetric orthogonal sequences can be obtained. For instance, a main class of symmetric orthogonal functions (MCSOF) has recently been introduced in [6]. Since we need the main properties of this class in the next sections let us restate them here.

2. Generation of MCSOF using Theorem 1.1

If the options

A(x) = x2(px2 + q) even; B(x) = x(rx2 + s) odd; C(x) = x2 > 0 even; (5) D(x) = 0 even; E(x) = −θ(s + (θ − 1)q) even; for p, q, r, s ∈ R; (−1)θ = −1,

(θ) n n and λn = − (n + (θ − 1)(1 − (−1) )/2)(r + (n − 1 + (θ − 1)(1 − (−1) )/2)p) are substituted in the main equation (3) the second order differential equation 2 2 + 00 + 2 + 0 + (θ) 2 − + − − − n  = x (px q)Φn (x) x(rx s)Φn(x) λn x θ(s (θ 1)q)(1 ( 1) )/2 Φn(x) 0, (6) will appear. According to [6], one of the basic solutions of this equation is a symmetric sequence of orthogonal functions in the form

[n/2]−1 n   + + − − + 1−(−1)n (θ) r s Y (2j 1 (1 ( 1) )θ) q s ( )(θ−1) S x = x 2 n p q 2j − 1 + n + 1 − −1 n − 1 2 p + r j=0 ( ( ( ) )(θ / ))

[n/2]   [n/2]−(k+1) n ! X [n/2] Y (2j − 1 + n + (1 − (−1) )(θ − 1/2)) p + r − × xn 2k, (7) k 2j + 1 + 1 − −1 n q + s k=0 j=0 ( ( ( ) )θ) which satisfies the recurrence relation         (θ) r s  1+(−1)n(θ−1) (θ) r s (θ) r s (θ) r s S + x = x S x + C S − x , (8) n 1 p q n p q n p q n 1 p q where r s C(θ) n p q

1 + (−1)n(θ − 1) pqn2 + (θ − 3 + 3(−1)n(1 − θ))pq + (1 + (−1)n(θ − 1))qr − (−1)nθps n + ((1 − θ)(p − r)q + ((θ − 3)p + r)s) (1 − (−1)n)/2 = . (8.1) (((−1)n(θ − 1) + 2n + θ − 2)p + r)(((−1)n(1 − θ) + 2n + θ − 4)p + r) As the recurrence relation (8) shows, MCSOF is reduced to a main class of symmetric orthogonal polynomials (MCSOP) if and only if θ = 1. In other words, MCSOP satisfies the relation         (1) r s (1) r s (1) r s (1) r s S + x = xS x + C S − x , (9) n 1 p q n p q n p q n 1 p q in which

r s pqn2 + ((r − 2p)q − (−1)nps) n + (r − 2p)s(1 − (−1)n)/2 C(1) = . (9.1) n p q (2pn + r − p)(2pn + r − 3p) M. Masjed-Jamei, M. Dehghan / Computers and Mathematics with Applications 56 (2008) 2941–2947 2943

See [7] for more details. The functions (7) also satisfy a generic orthogonality relation as Z α       r s (θ) r s (θ) r s = W x Sn x Sm x dx Nnδn,m, (10) −α p q p q p q where   Z − 2 + −  Z − 2 +  r s 2 (r 4p)x (s 2q) (r 2p)x s W x = x exp dx = exp dx , (10.1) p q x(px2 + q) x(px2 + q) denotes the corresponding weight function and [ ] 2 n/2  n n  Y (1) r + (1 − (−1) )θp, s + (1 − (−1) )θq N = C n i p, q i=1 Z α  n n  r + (1 − (−1) )θp, s + (1 − (−1) )θq × W x dx, (10.2) −α p, q shows the generic value of the norm square and finally α takes the standard values 1, ∞. (θ) Although we introduced four main sub-classes of Sn (x; p, q, r, s) in [6], further important sub-classes can still be found for the specific values of p, q, r, s and θ. In this paper, we introduce one of the mentioned samples, which firstly generalizes Fourier trigonometric sequences and is secondly orthogonal with respect to the same constant weight function on [0, π]. Here it may be interesting for the reader that there is also a special trigonometric sequence, which is orthogonal with respect to the normal distribution exp(−γ x2) on (−∞, ∞) though it is not a special case of MCSOF. For more details, see [8]. To introduce a classical generalization of Fourier trigonometric sequences, we should first (for convenience) suppose that in equation (6) p = −1 and q = 1. In this case, by the change of variable x = cos t, the modified differential equation (6) is transformed to − + 3 −   − − n  2 00 (r 1) cos t s cos t 0 (θ) 2 1 ( 1) cos tΦ (t) + Φ (t) + λ cos t − θ(s + θ − 1) Φn(t) = 0. (11) n sin t n n 2 Now if r + 1 = 0 and s = 0 in (11), a generalization of differential equation (1) is derived. Hence, by substituting the initial vector (p, q, r, s, θ) = (−1, 1, −1, 0, θ) in equation (6) and applying the change of variable x = cos t we get  −  00 n 2 θ(θ 1) n Φ (t) + n + (θ − 1)(1 − (−1) )/2 − (1 − (−1) )/2 Φn(t) = 0. (12) n cos2 t According to the previous comments, (12) is a generalization of (1) for θ = 1 and has a basic solution as

[n/2]−1 n   + + − − 1−(−1)n (θ) −1 0 Y 2j 1 (1 ( 1) )θ ( )(θ−1) Φn(t) = S cos t = (cos t) 2 n −1 1 2j + n + 1 − −1 n − 1 2 j=0 ( ( ) )(θ / )

[n/2]   [n/2]−(k+1) n ! X [n/2] Y 2j + n + (1 − (−1) )(θ − 1/2) − × (−1)k cosn 2k t. (13) k 2j + 1 + 1 − −1 n k=0 j=0 ( ( ) )θ But, this solution can be simplified to an easier form because for n = 2m we have

  m−1 + m   m−(k+1) + ! (θ) −1 0 Y j 1/2 X k m Y j m 2m−2k 1 S cos t = (−1) cos t = cos 2mt, (14) 2m −1 1 j + m k j + 1 2 22m−1 j=0 k=0 j=0 / while for n = 2m + 1, (13) becomes

  m−1 + + m   m−(k+1) + + ! (θ) −1 0 Y j θ 1/2 (θ−1) X k m Y j θ m 2m+1−2k S + cos t = cos t (−1) cos t, (15) 2m 1 −1 1 j + + m k j + + 1 2 j=0 θ k=0 j=0 θ / which has a complicated form yet. To simplify (15), let us first assume that

  m−(j+1) + + ! j m Y i θ m Aj = (−1) . (15.1) j i + + 1 2 i=0 θ / Consequently   + m (θ) −1 0 (θ 1/2)m (θ−1) X 2m+1−2j S + cos t = cos t Aj cos t, (15.2) 2m 1 −1 1 + m (θ )m j=0 2944 M. Masjed-Jamei, M. Dehghan / Computers and Mathematics with Applications 56 (2008) 2941–2947 where (a)m = a(a + 1)...(a + m − 1). On the other hand, since n   + 1 X 2n + 1 cos2n 1 x = cos(2n + 1 − 2k)x, (16) 22n k k=0 we have m m m k   ! X 2m+1−2j X 2k+1 X 1 X 2k + 1 Aj cos t = Am−k cos t = Am−k cos(2k + 1 − 2i)t 22k i j=0 k=0 k=0 i=0

m = X (m) + − Bk cos(2m 1 2k)t, (17) k=0 in which

k   k j   m−(j+1) !   (m) X Aj 2m + 1 − 2j X (−1) m Y i + θ + m 2m + 1 − 2j B = = . (17.1) k 22(m−j) k − j 22(m−j) j i + + 1 2 k − j j=0 j=0 i=0 θ /

(m) So, it is sufficient to compute Bk . For this purpose, we need the following identities

k 0(β)(−1) + 0(β + k) = 0(β)(β)k and 0(β − k) = ; β ∈ R; k ∈ Z , (18) (1 − β)k (m) where 0(β) denotes the gamma function [2]. After some calculations and applying the above identities, Bk is simplified as

k j (m) X 4 0(θ + 1/2)0(θ + 2m)0(2m + 2 − 2j) (−k)j(k − 2m − 1)j(−m)j(1/2 − θ − m)j B = . (19) k 22m0 + m 0 + m + 1 2 0 2m + 2 − k K! 1 − − 2m j! j=0 (θ ) (θ / ) ( ) ( θ )j

(m) As we observe, Bk is not in a hypergeometric form yet (see e.g. [9]) and one should apply the duplication Legendre formula [1]

22z−1  1  0(2z) = √ 0(z)0 z + , (20) π 2 for 0(2m + 2 − 2j) in (19) to finally get

k j (m) 2 0(θ + 1/2)0(θ + 2m)0(m + 3/2)m! X (−k)j(k − 2m − 1)j(1/2 − θ − m)j 1 B = √ k 0 + m 0 + m + 1 2 0 2m + 2 − k K! −m − 1 2 1 − − 2m j! π (θ ) (θ / ) ( ) j=0 ( / )j( θ )j   2 0(θ + 1/2)0(θ + 2m)0(m + 3/2)m! −k k − 2m − 1 1/2 − θ − m = √ 3F2 1 , (21) π 0(θ + m)0(θ + m + 1/2)0(2m + 2 − k)K! −m − 1/2 1 − θ − 2m

  k a b c P∞ (a)k(b)k(c)k x where 3F2 x = = denotes the of order (3, 2) [9]. Fortunately, the d e k 0 (d)k(e)k k! hypergeometric value appearing in (21) is a special case of Saalschutz theorem [10] that says if c is a negative integer and a + b + c + 1 = d + e then   a b c (d − a)|c|(d − b)|c| 3F2 1 = . (22) d e (d)|c|(d − a − b)|c| Therefore, after some calculations, (15.2) is finally simplified as   m   − − + (θ) −1 0 1 (θ−1) X 2m + 1 (θ 1)k( m 1/2)k S + cos t = cos t cos(2m + 1 − 2k)t. (23) 2m 1 −1 1 22m k 1 − − 2m −m − 1 2 k=0 ( θ )k( / )k

(θ) Corollary 1. The trigonometric sequence Φn (t) defined as  (θ) 1 Φ (t) = cos 2nt,  2n 22n−1 n   − − + (24) (θ) 1 (θ−1) X 2n + 1 (θ 1)k( n 1/2)k Φ + (t) = cos t cos(2n + 1 − 2k)t,  2n 1 22n k 1 − − 2n −n − 1 2 k=0 ( θ )k( / )k

(1) satisfies the differential equation (12). Clearly Φn (t) leads to the usual cosine sequence. M. Masjed-Jamei, M. Dehghan / Computers and Mathematics with Applications 56 (2008) 2941–2947 2945

To compute the norm square value of this sequence, one should refer to relations (10), (10.1) and (10.2). Hence, substituting (p, q, r, s, θ) = (−1, 1, −1, 0, θ) and α = 1 in (10) yields Z 1     Z π     1 (θ) −1 0 (θ) −1 0 (θ) −1 0 (θ) −1 0 √ S x S x dx = S cos t S cos t dt 2 n −1 1 m −1 1 n −1 1 m −1 1 −1 1 − x 0 [ ] 2 n/2  n n  Y (1) −1 + ((−1) − 1)θ, (1 − (−1) )θ = C i −1 1 i=1

Z 1  n n   −1 + ((−1) − 1)θ, (1 − (−1) )θ × W x dx δ . (25) −1 1 n,m −1 Now let n → 2n + 1 in (25). According to Corollary 1, the orthogonality (25) changes to

Z π 2n   Z 1 ! (θ) (θ) Y (1) −1 − 2θ, 2θ − Φ (t)Φ (t)dt = C x2θ (1 − x2) 1/2dx δ 2n+1 2m+1 i −1 1 n,m 0 i=1 −1 ! √ 0(θ + 1/2) Y2n (i + θ(1 − (−1)i))(−i + 1 − θ(1 − (−1)i)) = π δn,m, (26) 0 + 1 4 i + i + − 1 (θ ) i=1 ( θ)( θ ) if and only if θ + 1 > 0 and (−1)θ = −1. This result leads to the following final corollary:

(a) Corollary 2. The modified trigonometric sequence Cn (x) defined as

 (a) C (x) = cos 2nx,  2n n   − + (a) a  X 2n + 1 (a)k( n 1/2)k (27) C + (x) = cos x cos(2n + 1 − 2k)x,  2n 1 k −a − 2n −n − 1 2  k=0 ( )k( / )k satisfies the orthogonality relation

0 if i 6= j,  = = Z π π if i j 0, (a) (a) = = = Ci (x)Cj (x)dx π/2 if i j 2n, (28) 0  √ 0(a + 3/2 + n)0(a + 1 + n) 22n(2n)! π if i = j = 2n + 1,  0(a + 2 + 2n)0(a + 1 + 2n) if and only if a > −2 and (−1)a = 1. Note that by using (20) the last equality of (28) is equal to π/2 if a = 0 is considered.

(a) 3. Application of trigonometric sequence Cn (x) in functions expansion theory

Let f (x) be a periodic function satisfying Dirichlet conditions [1,2] and suppose for convenience that

(a) = a (a) C2n+1(x) (cos x)Dn (x). (29) According to section 9.4.‘‘Completeness of Eigenfunctions’’ of [1, pp. 523–538], Courant and Hilbert were probably the first persons who proved in their book, i.e. Methods of Mathematical physics (1953), that any sequence of the eigenfunctions of a { (a) }∞ Sturm–Liouville problem is a complete set of orthogonal functions. Hence, one can conclude that Cn (x) n=0 are a complete orthogonal set over [0, π] and the periodic function f (x) can be expanded in terms of them as

∞ ∞ ∞ 1 (a) X (a) 1 X a X (a) f (x) = b0C (x) + bkC (x) = b0 + b2k cos 2kx + (cos x) b2k+1D (x), (30) 2 0 k 2 k k=1 k=1 k=0 in which  2 Z π b2k = cos 2kxf (x)dx,  π  0 + + + + Z π  0(a 2 2k)0(a 1 2k) a (a) b + = √ f (x) cos xD (x)dx, 2k 1 2k k (30.1) π2 (2k)!0(a + 3/2 + k)0(a + 1 + k) 0  k k    (a) X X 2k + 1 (a)j(−k + 1/2)j D (x) = dk(a) cos(2k + 1 − 2j)x = cos(2k + 1 − 2j)x.  k j j −a − 2k −k − 1 2 j=0 j=0 ( )j( / )j 2946 M. Masjed-Jamei, M. Dehghan / Computers and Mathematics with Applications 56 (2008) 2941–2947

As (30.1) shows, b2k+1 can be simplified as

Z π k Z π (a) = a (a) = X k a + − Rk f (x) cos xDk (x)dx dj (a) f (x) cos x cos(2k 1 2j)xdx. (31) 0 j=0 0

One of the interesting choices for a in relation (31), by noting that a > −2 and (−1)a = 1, is when a is an even integer, because in this case

m 1 X 2m cos2m x = cos(2m − 2k)x; m ∈ N, (32) 22m−1 k k=0 and consequently

k m   Z π ! (2m) 1 X X 2m R = dk(2m) f (x) cos(2m − 2r)x cos(2k + 1 − 2j)xdx k 22m−1 j r j=0 r=0 0 k m π 1 X X 2m Z = dk(2m) f (x) cos(2m − 2r + 2k + 1 − 2j)xdx 22m j r j=0 r=0 0

Z π  ! + f (x) cos(2m − 2r − 2k − 1 + 2j)xdx . (33) 0

Let us present a practical example here. It can straightforwardly be verified that the usual cosine series of function f (x) = x2 over [0, π] is as follows

∞ ∞ ∞ π 2 X (−1)k π 2 X 1 X 1 x2 = + 4 cos kx = + cos 2kx − 4 cos(2k + 1)x. (34) 3 k2 3 k2 2k + 1 2 k=1 k=1 k=0 ( )

Now, to derive the generalized cosine series of type (30), if we suppose a = 2m then we get

k m    ! (2m) −π X X 2m 1 1 R = dk(2m) + . (35) k 22m−1 j r 2m − 2r + 2k + 1 − 2j 2 2m − 2r − 2k + 1 − 2j 2 j=0 r=0 ( ) ( ( ))

Hence, the corresponding odd coefficients take the form

+ + ! + ! k   − + (m) 2m+2 (2m 2k 1) (2m 2k) X 2k + 1 (2m)j( k 1/2)j (m) b + = −2 S (j, k), (36) 2k 1 4m + 2k + 1 ! 2k ! j −2m − 2k −k − 1 2 ( ) ( ) j=0 ( )j( / )j where

m X 2m  1 1  S(m)(j, k) = + . (36.1) r 2m − 2r + 2k + 1 − 2j 2 2m − 2r − 2k + 1 − 2j 2 r=0 ( ( )) ( ( ))

This gives the desired series as

2 ∞ ∞ 2 π X 1 2m X (m) (2m) x = + cos 2kx + (cos x) b + D (x), (37) 3 k2 2k 1 k k=1 k=0

= (0) = + (0) = − (0) = − + 2 which is a generalization of (34) for m 0 because Dk (x) cos(2k 1)x and b2k+1/2 2S (0, k) 4/(2k 1) . = (0) Of course, we should note that the identity (32) is not valid for m 0 and thus b2k+1/2 should be considered in (37) (not (0) = b2k+1), though for the rest of values of m (37) is valid. For instance, if m 1 then

+ + k   − + (1) (k 1)(2k 1) X 2k + 1 (2)j( k 1/2)j (1) b + = −16 S (j, k), (37.1) 2k 1 k + 2 2k + 5 j −2 − 2k −k − 1 2 ( )( ) j=0 ( )j( / )j 1 4 1 where S(1)(j, k) = + + . (37.1.1) (2k + 3 − 2j)2 (2k + 1 − 2j)2 (2k − 1 − 2j)2 M. Masjed-Jamei, M. Dehghan / Computers and Mathematics with Applications 56 (2008) 2941–2947 2947

Finally let us recall that there exists a similar case for a generalization of Fourier sine series. In other words, if one replaces the initial data (p, q, r, s, θ) = (−1, 1, −3, 0, θ) in the main orthogonality relation (10) for α = 1 as Z 1     Z π     p (θ) −3 0 (θ) −3 0 2 (θ) −3 0 (θ) −3 0 1 − x2S x S x dx = (sin t)S cos t S cos t dt n −1 1 m −1 1 n −1 1 m −1 1 −1 0 [ ] 2 n/2  n n  Y (1) −3 + ((−1) − 1)θ, (1 − (−1) )θ = C i −1 1 i=1

Z 1  n n   −3 + ((−1) − 1)θ, (1 − (−1) )θ × W x dx δ , (38) −1 1 n,m −1 then, by defining the trigonometric sequence   (a) n+1 (a+1) −3 0 S + (x) = 2 sin xS cos x , (39) n 1 n −1 1 one can again verify (for example) that   (a) 2n+1 (a+1) −3 0 S + (x) = 2 sin xS cos x = sin(2n + 1)x. (40) 2n 1 2n −1 1

References

[1] G. Arfken, Mathematical Methods for Physicists, Academic Press Inc, 1985. [2] A.F. Nikiforov, V.B. Uvarov, Special Functions of Mathematical Physics, Birkhauser, Basel, Boston, 1988. [3] M. Masjed-Jamei, A generalization of classical symmetric orthogonal functions using a symmetric generalization of Sturm–Liouville problems, Transforms Spec. Funct. 18 (2007) 871–883. [4] S.A. Chihara, An Introduction to Orthogonal Polynomials, Gordon and Breach, New York, 1978. [5] G. Szegö, Orthogonal Polynomials, 4th ed., Amer. Math. Soc., Providence, RI, 1975. [6] M. Masjed-Jamei, A basic class of symmetric orthogonal functions using the extended Sturm–Liouville theorem for symmetric functions, J. Comput. Appl. Math. 216 (2008) 128–143. [7] M. Masjed-Jamei, A basic class of symmetric orthogonal polynomials using the extended Sturm–Liouville theorem for symmetric functions, J. Math. Anal. Appl. 325 (2007) 753–775. [8] M. Masjed-Jamei, Biorthogonal exponential sequences with weight function exp(ax2 + ibx) on the real line and an orthogonal sequence of trigonometric functions, Proc. Amer. Math. Soc. 136 (2008) 409–417. [9] W. Koepf, Hypergeometric Summation, Vieweg, Braunschweig, Wiesbaden, 1988. [10] G.E. Andrews, R. Askey, R. Roy, Special Functions, in: Encyclopedia of Mathematics and its Applications, vol. 71, Cambridge University Press, Cambridge, 1999.