• Nie Znaleziono Wyników

1. Preliminaries. Basic notions of Algebraic Analysis We recall here the following notions and theorems (without proofs; cf.

N/A
N/A
Protected

Academic year: 2021

Share "1. Preliminaries. Basic notions of Algebraic Analysis We recall here the following notions and theorems (without proofs; cf."

Copied!
24
0
0

Pełen tekst

(1)

FOURIER-LIKE METHODS FOR EQUATIONS WITH SEPARABLE VARIABLES

Danuta Przeworska-Rolewicz

Institute of Mathematics Polish Academy of Sciences

Sniadeckich 8, 00–956 Warszawa 10, P.O. Box 21, Poland´ e-mail: rolewicz@impan.gov.pl

Abstract

It is well known that a power of a right invertible operator is again right invertible, as well as a polynomial in a right invertible opera- tor under appropriate assumptions. However, a linear combination of right invertible operators (in particular, their sum and/or differ- ence) in general is not right invertible. It will be shown how to solve equations with linear combinations of right invertible operators in com- mutative algebras using properties of logarithmic and antilogarithmic mappings. The used method is, in a sense, a kind of the variables separation method. We shall obtain also an analogue of the classical Fourier method for partial differential equations. Note that the results concerning the Fourier method are proved under weaker assumptions than these obtained in [6] (cf. also [7, 8, 11]).

Keywords: algebraic analysis, commutative algebra with unit, Leib- niz condition, logarithmic mapping, antilogarithmic mapping, right in- vertible operator, sine mapping, cosine mapping, initial value problem, boundary value problem, Fourier method.

2000 Mathematics Subject Classification: 47A25, 47C05, 47N20, 35A25.

1. Preliminaries. Basic notions of Algebraic Analysis We recall here the following notions and theorems (without proofs; cf.

[7, 8]). Denote by N, N

0

, R, C, Z, Q the sets of positive integers, nonnega-

tive integers, reals, complexes, integers and rational numbers, respectively,

(2)

and by F any field of scalars. If F is a field of numbers, then is denoted by F[t] the set of all polynomials in t with coefficients in F.

Let X be a linear space (in general, without any topology) over a field F of scalars of the characteristic zero.

• L(X) is the set of all linear operators with domains and ranges in X;

• dom A is the domain of an A ∈ L(X);

• ker A = {x ∈ dom A : Ax = 0} is the kernel of an A ∈ L(X);

• L

0

(X) = {A ∈ L(X) : dom A = X};

• I(X) is the set of all invertible elements in X.

An operator D ∈ L(X) is said to be right invertible if there is an operator R ∈ L

0

(X) such that RX ⊂ dom D and DR = I, where I denotes the identity operator. The operator R is called a right inverse of D. By R(X) we denote the set of all right invertible operators in L(X). Let D ∈ R(X).

Let R

D

⊂ L

0

(X) be the set of all right inverses for D, i.e., DR = I whenever R ∈ R

D

. We have dom D = RX ⊕ ker D, independently of the choice of an R ∈ R

D

. Elements of ker D are said to be constants, since by definition, Dz = 0 if and only if z ∈ ker D. The kernel of D is said to be the space of constants. We should point out that, in general, constants are different than scalars, since they are elements of the space X. If two right inverses commute with each other, then they are equal.

Clearly, if ker D 6= {0}, then the operator D is right invertible, but not invertible. Here the invertibility of an operator A ∈ L(X) means that the equation Ax = y has a unique solution for every y ∈ X. An element y ∈ dom D is said to be a primitive for an x ∈ X if y = Rx for an R ∈ R

D

. Indeed, by definition, x = DRx = Dy. Again, by definition, all x ∈ X have primitives. Let

F

D

= {F ∈ L

0

(X) : F

2

= F ; F X = ker D and ∃

R∈RD

F R = 0}.

Any F ∈ F

D

is said to be an initial operator for D corresponding to R. One can prove that any projection F

0

onto ker D is an initial operator for D corresponding to a right inverse R

0

= R − F

0

R independently of the choice of an R ∈ R

D

.

If two initial operators commute with each other, then they are equal.

Thus this theory is essentially noncommutative. An operator F is initial for D if and only if there is an R ∈ R

D

such that

(1.1) F = I − RD on dom D.

(3)

It is enough to know one right inverse in order to determine all right inverses and all initial operators. Note that a superposition of a finite number of right invertible operators is again a right invertible operator.

The equation Dx = y (y ∈ X) has the general solution x = Ry + z, where R ∈ R

D

is arbitrarily fixed and z ∈ ker D is arbitrary. However, if we take an initial condition: F x = x

0

, where F ∈ F

D

and x

0

∈ ker D, then this equation has a unique solution x = Rx + x

0

.

If T ∈ L(X) belongs to the set Λ(X) of all left invertible operators, then ker T = {0}. If D is invertible, i.e., D ∈ I(X) = R(X) ∩ Λ(X), then F

D

= {0} and R

D

= {D

−1

}.

If P (t) ∈ F[t] then all solutions of the equation P (D)x = y, y ∈ X, can be obtained by a decomposition of the rational function 1/P (t) into vulgar fractions. Write

v

F

A = {λ ∈ F \ {0} : I − λA is invertible} for A ∈ L(X).

Clearly, λ ∈ v

F

A if and only if 1/λ is a regular value of A. Let V (X) be the set of all Volterra operators, i.e.,

V (X) = {A ∈ L

0

(X) : A − λI is invertible for all λ ∈ F \ {0}}.

Then A ∈ V (X) if and only if v

F

A = F \ {0}.

If X is an algebra over F with a D ∈ L(X) such that x, y ∈ dom D implies xy, yx ∈ dom D, then we say that X is a D-algebra and we write D ∈ A(X). If X is a commutative algebra then

A

(X) is denoted by

A

(X).

Let D ∈

A

(X) and

(1.2) f

D

(x, y) = D(xy) − c

D

[xDy + (Dx)y] for x, y ∈ dom D, where c

D

is a scalar dependent on D only. Clearly, f

D

is a bilinear (i.e., linear in each variable) form which is symmetric when X is commutative, i.e. when D ∈

A

(X). This form is called a non-Leibniz component (cf. [7]).

If D ∈

A

(X) then the product rule in X can be written as follows:

D(xy) = c

D

[xDy + (Dx)y] + f

D

(x, y) for x, y ∈ dom D.

If D ∈

A

(X) and if D satisfies the Leibniz condition:

(1.3) D(xy) = xDy + (Dx)y for x, y ∈ dom D,

(4)

then X is said to be a Leibniz algebra. It means that in Leibniz algebras c

D

= 1 and f

D

= 0. The Leibniz condition implies that xy ∈ dom D whenever x, y ∈ dom D, i.e., Leibniz algebras are D-algebras. If X is a Leibniz algebra with unit e then e ∈ ker D, i.e., D is not left invertible. The set of commutative Leibniz D-algebras X with a D ∈ R(X) and with unit e ∈ dom D is denoted by L(D). Clearly, if X ∈ L(D) then e ∈ ker D.

Non-Leibniz components for powers of D ∈

A

(X) are determined by recurrence formulae (cf. [7, 8]).

Suppose that D ∈

A

(X) and λ 6= 0 is an arbitrarily fixed scalar. Then λD ∈

A

(X) and c

λD

= c

D

, f

λD

= λf

D

.

If D

1

, D

2

A

(X), the superposition D = D

1

D

2

exists and D

1

D

2

∈ A(X), then

(1.4) c

D1D2

= c

D1

c

D2

and for x, y ∈ dom D = dom D

1

∩ D

2

f

D1D2

(x, y) = f

D1

(x, y) + D

1

f

D2

(x, y) + +c

D1

c

D2

[(D

1

x)D

2

y + (D

2

x)D

1

y].

For higher powers of D in Leibniz algebras, by an easy induction from For- mulae (1.4) and the Leibniz condition, we obtain the Leibniz formula:

(1.5) D

n

(xy) = X

n k=0

 n k



(D

k

x)D

n−k

y for x, y ∈ dom D

n

(n ∈ N).

Let X ∈

A

(X). We denote by M (X) the set of all multiplicative mappings (not necessarily linear) with domains and ranges in X:

M (X) = {A : A(xy) = (Ax)(Ay) whenever x, y ∈ dom A ⊂ X}.

Suppose that X ∈

A

(X) and D ∈ R(X). An initial operator F for D is said to be almost averaging if F (zx) = zF x whenever z ∈ ker D, x ∈ X.

Clearly, every multiplicative operator F ∈ F

D

is almost averaging for F (zx) = (F z)(F x) = zF x if z ∈ ker D, x ∈ X, but not conversely (cf.

[7]). If X is a D-algebra such that dim ker D=1 and e ∈ ker D then all initial operators for D are almost averaging.

Suppose that D ∈

A

(X). Let a multifunction Ω : dom D −→ 2

dom D

be defined as follows:

Ωu = {x ∈ dom D : Du = uDx} for u ∈ dom D \ {0}.

(5)

The equation

Du = uDx for (u, x) ∈ graph Ω is said to be the basic equation. Clearly,

−1

x = {u ∈ dom D : Du = uDx} for x ∈ dom D.

The multifunction Ω is well-defined and dom Ω ⊃ ker D \ {0}.

Suppose that (u, x) ∈ graph Ω, L is a selector of Ω and E is a selector of Ω

−1

. By definitions, Lu ∈ dom Ω

−1

, Ex ∈ dom Ω and the following equations are satisfied:

Du = uDLu, DEx = (Ex)Dx.

Any invertible selector L of Ω is said to be a logarithmic mapping and its inverse E = L

−1

is said to be an antilogarithmic mapping. By G[Ω] we denote the set of all pairs (L, E), where L is an invertible selector of Ω and E = L

−1

. For any (u, x) ∈ dom Ω and (L, E) ∈ G[Ω] elements Lu, Ex are said to be logarithm of u and antilogarithm of x, respectively. The multifunction Ω is examined in [8] (also for noncommutative algebras).

Clearly, by definition, for all (L, E) ∈ G[Ω], (u, x) ∈ graph Ω we have (1.6) ELu = u, LEx = x; DEx = (Ex)Dx, Du = uDLu.

A logarithm of zero is not defined. If (L, E) ∈ G[Ω] then L(ker D \ {0}) ⊂ ker D, E(ker D) ⊂ ker D. In particular, E(0) ∈ ker D.

If D ∈ R(X) then logarithms and antilogarithms are uniquely deter- mined up to a constant. Moreover, if F ∈ F

D

then F E = EF , F L = LF (cf. [15]).

Let D ∈

A

(X) and let (L, E) ∈ G[Ω]. A logarithmic mapping L is said to be of the exponential type if L(uv) = Lu + Lv for u, v ∈ dom Ω. If L is of the exponential type then E(x + y) = (Ex)(Ey) for x, y ∈ dom Ω

−1

. We have proved that a logarithmic mapping L is of the exponential type if and only if X is a Leibniz commutative algebra (cf. [8]). Moreover, Le = 0, i.e., E(0) = e. In Leibniz commutative algebras with D ∈ R(X) a necessary and sufficient conditions for u ∈ dom Ω is that u ∈ I(X) (cf. [8]).

By Lg(D) we denote the class of these commutative algebras with D ∈

R(X) and with unit e ∈ dom Ω for which there exist invertible selectors

of Ω, i.e., there exist (L, E) ∈ G[Ω]. By

L

(D) we denote the class of these

commutative Leibniz algebras with unit e ∈ dom Ω for which there exist

(6)

invertible selectors of Ω. By these definitions, X ∈ Lg(D) is a Leibniz algebra if and only if X ∈

L

(D) and D ∈ R(X). This class we shall denote by

L

(D). It means that

L

(D) is the class of these commutative Leibniz algebras with D ∈ R(X) and with unit e ∈ dom Ω for which there exist invertible selectors of Ω, i.e., there exist (L, E) ∈ G[Ω].

In the same manner we define logarithmic and antilogarithmic mappings of higher order. Namely, let n ∈ N be arbitrarily fixed. Suppose that D ∈

A

(X). Let a multifunction Ω

n

: dom D

n

−→ 2

dom Dn

be defined as follows:

(1.7) Ω

n

u = {x ∈ dom D

n

: D

n

u = uD

n

x} for u ∈ dom D

n

.

Any invertible selector L

n

of Ω

n

is said to be a logarithmic mapping of the order n and its inverse E

n

= L

−1n

is said to be an antilogarithmic mapping of the order n. By G[Ω

n

] we denote the set of all pairs (L

n

, E

n

), where L

n

is an invertible selector of Ω

n

and E

n

= L

−1n

. For any (u, x) ∈ dom Ω

n

and (L

n

, E

n

) ∈ G[Ω

n

] elements L

n

u, E

n

x are said to be logarithm of the order n of u and antilogarithm of the order n of x, respectively. The multifunctions Ω

n

and relations between them are examined in [8]. Clearly, if X ∈ Lg(D) then X ∈ Lg(D

n

) for all n ∈ N.

If ker D = {0}, then either X is not a Leibniz algebra or X has no unit (cf. [8]). Thus, by our definition, if X ∈

L

(D), then ker D 6= {0}, i.e., the operator D is right invertible but not invertible.

2. Linear combinations of right invertible operators We begin with

Proposition 2.1. Suppose that n, r

1

, . . . , r

n

∈ N,

(2.1) X ∈

\

n j=1

L

(D

j

),

D = X

n j=1

α

j

D

rjj

, α

j

∈ X (j = 1, . . . , n), dom D =

\

n j=1

dom D

jrj

6= ∅, (2.2)

(L

(j)rj

, E

r(j)j

) ∈ G[Ω

(j)rj

], where Ω

(j)rj

is induced by D

rjj

(j = 1, . . . , n),

(7)

(2.3) x = Y

n k=1

u

k

, where u

k

∈ ker D

k

∩ I(X) (k = 1, . . . , n).

Then

(2.4) Dx = ax, where a =

X

n j=1

α

j

a

j

,

(2.5) a

j

= D

jrj

L

(j)rj

e u

j

and u e

j

= Y

n k=1, k6=j

u

k

(j = 1, . . . , n),

i.e., u e

j

, u

j

u e

j

= x ∈ I(X) (j = 1, . . . , n).

P roof. Since, by our assumptions, the operators D

1

,. . . ,D

n

satisfy the Leibniz condition and D

1

u

1

= ... D

n

u

n

= 0, from the Leibniz Formula (1.5) we get

D

mj

x = D

jm

 Y

n k=1

u

k



= X

m

l=0

 m l



(D

jl

u

j

)(D

jm−l

e u

j

) = u

j

D

jm

e u

j

for j = 1, . . . , n. Thus Dx =

 X

n j=1

α

j

D

rjj

 Y

n k=1

u

k

= X

n j=1

α

j

u

j

D

rjj

Y

n k=1, k6=j

u

k

=

 Y

n k=1

u

k

 X

n j=1

α

j

 Y

n k=1, k6=j

u

−1k

 D

rjj

 Y

n k=1, k6=j

u

k



= x X

n j=1

α

j

u e

−1j

D

jrj

u e

j

= x X

n j=1

α

j

D

jrj

L

(j)rj

e u

j

= x X

n j=1

α

j

a

j

= xa.

Proposition 2.2. Suppose that all assumptions of Proposition 2.1 are satis- fied. Then there are R

j

∈ R

Dj

such that R

jrj

a

j

∈ dom (Ω

(j)rj

)

−1

(j = 1, . . . , n) and

(2.6) u e

j

= E

r(j)j

(R

rjj

a

j

) (j = 1, . . . , n).

(8)

P roof. By our assumptions, a

j

= D

jrj

L

(j)rj

e u

j

for j = 1, . . . , n. Hence there are R

j

∈ R

Dj

such that L

(j)rj

u e

j

= R

rjj

a

j

(cf. [8]), i.e., e u

j

= E

r(j)j

L

(j)rj

e u

j

= E

r(j)j

(R

rjj

a

j

).

Proposition 2.3. Suppose that all assumptions of Proposition 2.1 are sat- isfied and r

1

= . . . = r

n

= 1. Then the operator D defined by (2.2) satisfies the Leibniz condition.

P roof. Let x, y ∈ dom D. Clearly, x, y ∈ dom D whenever x, y ∈ T

n

j=1

dom D

j

. Since D

1

, . . . , D

n

satisfy the Leibniz condition, we get D(xy) =

X

n j=1

α

j

D

j

(xy) = X

n j=1

α

j

(xD

j

y + yD

j

x) =

= x X

n j=1

α

j

D

j

y + y X

n j=1

α

j

D

j

x = xDy + yDx.

Proposition 2.4. Suppose that all assumptions of Proposition 2.1 are sat- isfied. Let

(2.7) U

n

=

 Y

n k=1

u

k

: u

k

∈ ker D

k

∩ I(X) (k = 1, . . . , n)



(n ∈ N).

Then selectors L of the multifunction Ω induced by D satisfy the equality DLx = a for x ∈ U

n

.

P roof. By Equation (2.4), we have Dx = ax, where x ∈ I(X). Thus, by definition, DLx = x

−1

Dx = a for any selector L of Ω.

Propositions 2.3 and 2.4 imply

Corollary 2.1. Suppose that all assumptions of Proposition 2.1 are satis- fied. If r

1

= . . . = r

n

= 1 and U

n

∈ Lg(D), then U

n

L

(D).

Proposition 2.5. Suppose that all assumptions of Proposition 2.1 are sat- isfied and a = 0. Then there are R

j

∈ R

Dj

such that R

jrj

a

j

∈ dom (Ω

(j)rj

)

−1

(j = 1, . . . , n) and

(2.8) x = 1

n X

n j=1

u

j

u e

j

= 1 n

X

n j=1

u

j

E

r(j)j

(R

rjj

a

j

) ∈ ker D,

(9)

where u

j

∈ ker D

j

∩ I(X) (j = 1, . . . , n).

P roof. By our assumptions and Proposition 2.1, Dx = ax = 0 and x = u

j

u e

j

(j = 1, . . . , n). Hence x =

n1

P

n

j=1

u

j

e u

j

. This, and Proposition 2.2 together imply (2.8).

Proposition 2.6. Suppose that all assumptions of Proposition 2.1 are sat- isfied and a ∈ I(X). Then the equation

(2.9) Dx = y, y ∈ X

has a solution

(2.10) x = y

 X

n j=1

α

j

D

rjj

L

(j)rj

u e

j



−1

.

P roof. By our assumptions, ax = Dx = y. Since a ∈ I(X), we get x = a

−1

y. Propositions 2.1 and 2.2 together imply that

x = a

−1

y = y

 X

n j=1

α

j

a

j



−1

= y

 X

n j=1

α

j

D

rjj

L

(j)rj

u e

j



−1

.

Corollary 2.2. Suppose that all assumptions of Proposition 2.1 are sat- isfied, r

1

= . . . = r

n

= 1 and a ∈ I(X). Then the equation (2.9) has a solution

(2.11) x = y

 X

n j=1

α

j

D

j

X

n k=1, k6=j

L

(j)1

u

k



−1

.

P roof. Proposition 2.6 and the Leibniz condition together imply that x = y

 X

n j=1

α

j

D

j

L

(j)1

Y

n k=1, k6=j

u

k



−1

= y

 X

n j=1

α

j

D

j

X

n k=1, k6=j

L

(j)1

u

k



−1

.

Some more generalized approaches to problems with linear combinations of

right invertible operators of order one on vectors fields and magnifolds have

been given by Virsik [17] and Multarzy´ nski [3, 4].

(10)

3. Trigonometric elements and mappings

We shall show now an approach to the trigonometric identity in Leibniz D- algebras with unit e (but not necessarily with logarithms). Clearly, without additional assumptions we cannot expect too much.

Proposition 3.1. Suppose that X ∈ L(D), x ∈ dom D

2

and x, Dx are not zero divisors. If x

2

+ (Dx)

2

= e, then

(3.1) αx + βDx ∈ ker (D

2

+ I) whenever α, β ∈ F.

P roof. Let y = −Dx. Then Dy = −D

2

x and

0 = De = D[x

2

+(Dx)

2

] = 2xDx+2(Dx)D

2

x = 2(Dx)(x+D

2

x) = 2y(x−Dy).

Since y = −Dx is not a zero divisor, we have x − 2y = 0. Hence Dy = x and y = −Dx = −D

2

y, which implies y ∈ ker (D

2

+ I). On the other hand, x = Dy = −D

2

x, which implies x ∈ ker (D

2

+ I).

Proposition 3.2. Suppose that all assumptions of Proposition 3.1 are sat- isfied. If Condition (3.1) holds for x and Dx and u = x

2

+ (Dx)

2

, then u ∈ ker D.

P roof. Let u = x

2

+ (Dx)

2

. Then Du = 2xDx+2(Dx)D

2

x = 2(Dx)(x + Dx) = 2(Dx)(D

2

+ I)x = 0, which implies u ∈ ker D.

Corollary 3.1. Suppose that all assumptions of Proposition 3.1 are sat- isfied, Condition (3.1) holds for x and Dx, F ∈ F

D

∩ M (X), F x = e, F Dx = 0 and u = x

2

+ (Dx)

2

. Then u = e, i.e., x

2

+ (Dx) = e.

P roof. Since F is a multiplicative initial operator and F x = e, F Dx = 0, we find u = F [x

2

+ (Dx)

2

] = (F x)

2

+ (F Dx)

2

= e

2

+ 0 = e.

Proposition 3.3. Suppose that all assumptions of Proposition 3.1 and Condition (3.1) are satisfied, F ∈ F

D

∩ M (X) and F x = e. Then F Dx = 0.

P roof. By our assumptions, e = F e = F [x

2

+ (Dx)

2

]=(F x)

2

+ (F Dx)

2

=

e + (F Dx)

2

, which implies (F Dx)

2

= 0. Hence F Dx = 0.

(11)

Proposition 3.4. Suppose that all assumptions of Proposition 3.1 are sat- isfied. If x

±

∈ ker (D ± iI) and x =

12

(x

+

+ x

), y =

2i1

(x

+

− x

), then

(i) x, y ∈ ker (D

2

+ I), Dx = −y, Dy = x and

12

(x ± y) ∈ ker (D

2

∓ iI);

(ii) x

2

+ y

2

= x

+

x

∈ ker D.

P roof. Points (i) is proved by checking. In order to prove (ii), observe that, by the Leibniz condition and our assumptions,

D(x

+

x

) = x

+

Dx

+ x

Dx

+

= ix

+

x

− ix

+

x

= 0.

Observe that x

±

are eigenvectors of the operator D corresponding to the eigenvalues ∓i, respectively.

Here and in the sequel we assume that F is an algebraically closed field of scalars. For instance, F = C. The following results are slightly stronger (with some proofs slightly simpler than in [7]):

Definition 3.1. Let X be a linear space over F. If λ ∈ F is an eigenvalue of an operator D ∈ R(X), then every eigenvector x

λ

corresponding to λ is said to be an exponential element (shortly: an exponential).  This means that x

λ

is an exponential if and only if x

λ

6= 0 and x

λ

∈ ker(D − λI).

Proposition 3.5. Let X be a linear space (over F). Suppose that D ∈ R(X). If 0 6= x

λ

∈ ker(I − λR) for an R ∈ R

D

and a λ ∈ F, then x

λ

∈ ker(D − λI), i.e., x

λ

is an exponential.

P roof. By our assumption, (D − λI)x

λ

=(D − λDR)x

λ

= D(I − λR)x

λ

=0.

By an easy induction we get

Proposition 3.6. Suppose that X is a linear space (over F), D ∈ R(X) and {λ

n

} ⊂ F is a sequence of eigenvalues such that λ

i

6= λ

j

for i 6= j. Then for an arbitrary n ∈ N the exponentials x

λ1

, . . . , x

λn

are linearly independent.

Proposition 3.7. Suppose that X is a linear space (over F), D ∈ R(X),

F is an initial operator for D corresponding to an R ∈ R

D

and x

λ

is an

exponential. Then x

λ

is an eigenvector for R corresponding to the eigenvalue

1/λ if and only if F x

λ

= 0, i.e., if R is not a Volterra operator.

(12)

P roof. Sufficiency. Since Dx

λ

= λx

λ

and F x

λ

= 0, we get x

λ

= x

λ

− F x

λ

= (I − F )x

λ

= RDx

λ

= λRx

λ

. Hence x

λ

∈ ker(I − λR). Since x

λ

6= 0, we conclude that x

λ

is an eigenvector for R corresponding to 1/λ.

Necessity. Suppose that 1/λ is an eigenvalue of R and the corresponding eigenvector x

λ

is an exponential. Then F x

λ

= (I − RD)x

λ

= (I − λR)x

λ

=

−λ(R −

λ1

I)x

λ

= 0.

Theorem 3.1. Suppose that X is a linear space (over F), D ∈ R(X), ker D 6= {0}, R ∈ R

D

and λ ∈ v

F

R. Then

(i) λ is an eigenvalue of D and the corresponding exponential is (3.2) x

λ

= e

λ

(z), where e

λ

= (I − λR)

−1

, z ∈ ker D;

whenever e

λ

= (I − λR)

−1

exists, is said to be an exponential operator;

(ii) the dimension of the eigenspace X

λ

corresponding to the eigenvalue λ is equal to the dimension of the space of constants, i.e., dim X

λ

= dim ker D 6= 0;

(iii) if λ 6= 0, then there exist non-trivial exponentials: e

λ

(z) 6= 0;

(iv) exponentials are uniquely determined by their initial values, i.e., if F is an initial operator for D corresponding to R, then F [e

λ

(z)] = z;

(v) if R is a Volterra operator, then every λ ∈ F is an eigenvalue of D, i.e., for every λ ∈ F there exist exponentials.

P roof. (i) By definition, (I − λR)e

λ

(z) = (I − λR)(I − λR)

−1

z = z, where z ∈ ker D. Thus e

λ

(z) = z + λRe

λ

(z), which implies De

λ

(z) = Dz + λDRe

λ

(z)=λe

λ

(z).

(ii) Since by our assumptions, the operator e

λ

= I −λR is invertible, dim X

λ

=dim {e

λ

(z) : z ∈ ker D}=dim {(I − λR)

−1

z : ker D}=dim ker D 6= 0.

(iii) If λ 6= 0 and e

λ

(z) = (I − λR)

−1

z = 0 then z = (I − λR)e

λ

(z) = 0, This contradicts our assumption that ker D 6= {0}.

(iv) By definitions and (i), we have F e

λ

(z) = (I − RD)e

λ

(z) = (I − λR) e

λ

(z) = z.

(v) If R ∈ V (X) then v

F

R = F \ {0}. Clearly, for λ = 0 the operator I − λR is also invertible. Hence, by (i), every scalar λ is an eigenvalue of D.

Definition 3.2. Let F = C. Suppose that X is a linear space (over C,

D ∈ R(X), ker D 6= {0} and R ∈ R

D

∩ V (X). Then the operators

(13)

(3.3) c

λ

= 1

2 (e

λi

+ e

−λi

), s

λ

= 1

2i (e

λi

− e

−λi

) (λ ∈ R)

are said to be cosine and sine operators, respectively (or: trigonometric operators). Elements c

λ

(z), s

λ

(z), where z ∈ ker D, are said to be cosine and sine elements, respectively (or: trigonometric elements).  Theorem 3.2. Suppose that all assumptions of Definition 3.2 are satisfied.

Then

(3.4) c

λ

= (I + λ

2

R

2

)

−1

, s

λ

= λR(I + λ

2

R

2

)

−1

(λ ∈ R)

(3.5) Dc

λ

= −λs

λ

, Ds

λ

= λc

λ

(λ ∈ R)

(3.6) c

0

(z) = z, s

0

(z) = 0, F s

λ

(z) = 0 f or z ∈ ker D, λ ∈ R.

Moreover, whenever z ∈ ker D, λ ∈ R, the element c

λ

(z) is even with respect to λ and the element s

λ

is odd with respect to λ.

P roof. By the first Formula of (3.4), for λ ∈ R we get c

λ

= 1

2

 (I − λiR)

−1

+ (I + λiR)

−1



= 1

2 (I − λiR)

−1

(I + λiR)

−1

(I + λiR + I − λiR)

= 1

2 (I + λ

2

R

2

)

−1

2I = (I + λ

2

R

2

)

−1

. A similar proof for s

λ

. By definitions, if λ ∈ R, then

Dc

λ

= 1

2 D(e

λi

+ e

−λi

) = 1

2 (λie

λi

+ λie

−λi

)

= 1

2 λi(e

λi

+ e

−λi

) = − λ

2i (e

λi

+ e

−λi

) = −λs

λ

.

Since DR = I, we have Ds

λ

= λDR(I + λ

2

R

2

)

−1

= λ(I + λ

2

R

2

)

−1

= λc

λ

.

(14)

Let z ∈ ker D. Let λ = 0. Then c

0

(z) = z, s

0

(z) = 0. Since F R = 0, for every λ ∈ R we have F s

λ

(z) = λF R(I + λ

2

R

2

)

−1

= 0. Let z ∈ ker D. Then

c

−λ

(z) = [I + (−λ)

2

R

2

)

−1

](z) = (I + λ

2

R

2

)

−1

z = c

λ

(z);

s

−λ

(z) = −λR[I + (−λ)

2

R

2

)

−1

](z) = −λR(I + λ

2

R

2

)

−1

z = −s

λ

(z).

Consider now trigonometric elements in algebras. It is easy to verify Proposition 3.6. Suppose that D ∈

A

(X) ∩ R(X), ker D 6= {0} and R ∈ R

D

∩ V (X). Then

(3.6) [c

λ

(z)]

2

+ [s

λ

(z)]

2

= e

λi

(z)e

−λi

(z) f or all z ∈ ker D, λ ∈ R.

(3.7) D[e

λi

(z)e

−λi

(z)] = c

D

z+f

D

(e

λi

(z), e

−λiz

) f or all z ∈ ker D, λ ∈ R.

Corollary 3.2. Suppose that X is a Leibniz D-algebra, ker D 6= {0} and R ∈ R

D

∩ V (X). Then

(3.8) D[e

λi

(z)e

−λi

(z)] = z f or all z ∈ ker D, λ ∈ R,

P roof. Since X is a Leibniz D-algebra, we have c

D

=1 and f

D

= 0. Hence Formulae (3.6) and (3.7) imply (3.8).

An immediate consequence of Corollary 3.2 is

Corollary 3.3. Suppose that X is a Leibniz D-algebra, ker D 6= {0} and R ∈ R

D

∩ V (X). Then the Trigonometric Identity holds, i.e.,

(3.9) [c

λ

(z)]

2

+ [s

λ

(z)]

2

= z f or all z ∈ ker D, λ ∈ R.

Proposition 3.7 (cf. [8]). Suppose that X ∈ Lg(D), λg = Re ∈ dom Ω

−1

for every R ∈ R

D

and λ ∈ v

F

R. Then there are (L, E) ∈ G[Ω] such that

E(λg) = (I − λR)

−1

z = e

λ

z ∈ ker(D − λI) f or all z ∈ ker D.

P roof. Let R ∈ R

D

be fixed. Elements of the form u = e

λ

z = (I −λR)

−1

z

are well-defined for all z ∈ ker D and (D − λI)u = D(I − λR)u = Dz = 0.

(15)

Moreover, Du = λu = uλe = uλDRe = uD(λg), which implies that λg ∈ dom Ω

−1

and there are (L, E) ∈ G[Ω] such that e

λ

= u = E(λg).

Definition 3.3 (cf. [8]). Suppose that F = C, X ∈ Lg(D) and E

1

= dom Ω

−1

is symmetric, i.e., −x ∈ E

1

whenever x ∈ E

1

. Let (L, E) ∈ G[Ω].

Write

(3.10) Cx = 1

2 [E(ix) + E(−ix)], Sx = 1

2i [E(ix) − E(−ix)] for ix ∈ E

1

. The mappings C and S are said to be cosine and sine mappings or trigono- metric mappings. Elements Cx and Sx are said to be cosine and sine

elements or trigonometric elements. 

Clearly, trigonometric mappings and elements have such properties as the classical cosine and sine functions. Namely, we have (proofs can be found in [8]):

Proposition 3.8 (cf. [8]). Suppose that all assumptions of Definition 3.3 are satisfied. Let (L, E) ∈ G[Ω]. Then trigonometric mappings C and S are well-defined for all ix ∈ E

1

and have the following properties:

(i) The de Moivre formulae hold:

E(ix) = Cx + iSx, E(−ix) = Cx − iSx f or ix ∈ E

1

. In particular, if X is a commutative Leibniz algebra then

(3.11) (Cx + iSx)

n

= C(nx) + iS(nx) f or ix ∈ E

1

and n ∈ N;

(ii) C and S are even and odd functions of their argument, respectively, i.e., C(−x) = Cx, S(−x) = −Sx for ix ∈ E

1

and C(0) = z ∈ ker D \ {0}, S(0) = 0. In particular, for all ix ∈ E

1

(3.12) (Cx)

2

+ (Sx)

2

= 1

2 [E(ix)E(−ix) + E(ix)E(ix)].

(iii) The mappings C

0

, S

0

defined as follows: C

0

x = C(x+z), S

0

x = S(x+z) for ix ∈ E

0

, z ∈ ker D also satisfy assertions (i)–(ii).

(iv) For all ix ∈ dom Ω

−1

(16)

(3.13) (Cx)

2

+ (Sx)

2

= E(ix)(E(−ix);

(3.14) DCx = −(Sx)Dx, DSx = (Cx)Dx.

Corollary 3.4 (cf. [8]). Suppose that all assumptions of Proposition 3.8 are satisfied and X is a Leibniz D-algebra with unit e. Then the Trigonometric Identity holds, i.e.,

(3.15) (Cx)

2

+ (Sx)

2

= e whenever ix ∈ E

1

.

The following question arises: Do non-Leibniz algebras with the Trigono- metric Identity (3.15) exist? The answer to this question is negative, i.e., non-Leibniz algebras with the Trigonometric Identity (3.15) do not exist (cf.

[11]). In other words: The Leibniz condition is necessary and sufficient for the Trigonometric Identity to hold.

In order to apply trigonometric mappings, we shall make use of the following condition:

[C]

n

F = C, n ∈ N is arbitrarily fixed, X ∈ Lg(D

n

), Ω

1

= Ω and dom Ω

−1n

is symmetric, i.e., −x ∈ dom Ω

n

whenever x ∈ dom Ω

n

. Suppose now that Condition [C]

2

holds and X ∈

L

(D). Suppose that λ ∈ C, R ∈ R

D

, g = Re and λig ∈ dom Ω

1

. If (L

1

, E

1

) ∈ G[Ω

1

] and (L

2

, E

2

) ∈ G[Ω

2

] then

(3.16)

ker(D

2

+ λ

2

I) = {z

1

E(λig) + z

2

E(−λig) : z

1

, z

2

∈ ker D}

= {zC(λg) + e zS(λg) : z, e z ∈ ker D}

= {(z

00

g + z

0

)E

2

( λ

2

g

2

2 ) : z

0

, z

00

∈ ker D}.

The assumption that λi,−λi ∈ v

C

R ensures that λig, −λig ∈ dom Ω

−11

. In

this case, −λ

2

∈ v

C

R

2

.

(17)

4. Fourier-like problems for right invertible operators We will apply properties of trigonomeric mappings and elements in order to find non-trivial solutions of some homogeneous initial and boundary value problems for difference of two right invertible operators (of the first and second order).

Proposition 4.1. Let X ∈ L(D

i

), ker D

i

6= {0}, i = 1, 2, ker D

1

∩ ker D

2

= Ce = {λe}

λ∈C

. Suppose that x = uv, where u ∈ ker D

2

, v ∈ ker D

1

. Then

(4.1) (D

2

− D

21

)x = u(D

2

+ λ

2

I)v − v(D

12

+ λ

2

I)v f or all λ ∈ C {0}.

P roof. By our assumptions, D

2

u = 0, D

1

v = 0 and both operators D

1

, D

2

satisfy the Leibniz condition. This and Leibniz Formula (1.5) together imply that

(D

2

− D

12

)x = (D

2

− D

12

)(uv) = D

2

(uv) − D

12

(uv)

= uD

2

v + vD

2

u − uD

12

v − 2(D

1

u)(D

1

v) − vD

21

u = uD

2

v − vD

12

u

= u(D

2

v + λ

2

v) − λ

2

uv − vD

1

u = u(D

2

+ λ

2

I)v − v(D

12

+ λ

2

I)u.

Proposition 4.1 immediately implies

Corollary 4.1. Let X ∈ L(D

i

), ker D

i

6= {0}, i = 1, 2, ker D

1

∩ ker D

2

= Ce = {λe}

λ∈C

. Suppose that x = uv, where u ∈ ker D

2

, v ∈ ker D

1

. Then (D

2

− D

21

)x = 0 if and only if u(D

2

+ λ

2

I)v − v(D

12

+ λ

2

I)v = 0 for all λ ∈ C {0}.

Corollary 4.2. Let X ∈ L(D

i

), ker D

i

6= {0}, i = 1, 2, ker D

1

∩ ker D

2

= Ce = {λe}

λ∈C

and x = uv, where u ∈ ker D

2

, v ∈ ker D

1

. Then (D

2

− D

21

)x = 0 if and only if u ∈ I(ker D

2

), v ∈ I(ker D

1

) and there is a λ ∈ C \ {0} such that

(4.2) u

−1

D

12

u = vD

2

v = −λ

2

e.

P roof. Equalities (4.2) hold if and only if (D

21

2

I)u = 0, (D

2

2

I)v = 0.

This, and Corollary 4.1 together imply that (D

2

− D

12

)x = 0 if and only if

Equalities (4.2) hold.

(18)

Theorem 4.1. Suppose that X ∈ L(D

i

), ker D

i

6= {0}, i = 1, 2, ker D

1

∩ ker D

2

= Ce = {λe}

λ∈C

and almost averaging F

0

, F

1

∈ F

D1

, F

2

∈ F

D2

correspond to R

0

, R

1

∈ R

D1

, R

2

∈ R

D2

, respectively. Suppose, moreover, that x = uv, where u ∈ I(ker D

2

), v ∈ I(ker D

1

) and there is a λ ∈ v

C

R

0

such that Equalities (4.2) hold and a u such that F

1

u=0. Then the homogeneous initial value problem

(4.3) (D

2

− D

12

)x = 0,

with the homogeneous boundary condition

(4.4) F

0

x = 0, F

1

x = 0

and with the homogeneous initial condition

(4.5) F

2

x = 0

is ill posed, since it has a non-trivial solution x = uv, where v is an eigen- vector of R

2

corresponding to the eigenvalue -λ

2

.

P roof. By Corollary 4.3, elements u, v are invertible by our assumption, hence they are not zero divisors and x = uv is a non-trivial solution of the equation (D

2

− D

12

)x = 0. Since λ ∈ v

C

R

0

, Equalities (4.2) imply that v ∈ ker (D

2

+ λ

2

I) = ker D

2

(I + λ

2

I)R

2

, i.e., v = −λ

−2

R

2

, u ∈ ker (D

12

+ λ

2

I).

Since F

0

R

0

= 0, we have F

0

u = 0. Since u ∈ ker D

2

, v ∈ ker D

1

and initial operators F

0

, F

1

, F

2

are almost averaging, we find F

0

x = F

0

(uv) = vF

0

u = 0, F

1

(uv) = vF

1

u = 0, F

2

x = F

2

(uv) = uF

2

v = u(−λ−2)F

2

R

2

v = 0 (for F

2

R

2

= 0).

Theorems 4.1 and 3.2 together imply

Corollary 4.3. Suppose that X ∈ L(D

i

), ker D

i

6= {0}, i = 1, 2, ker D

1

∩ ker D

2

= Ce = {λe}

λ∈C

and almost averaging F

0

, F

1

∈ F

D1

, F

2

∈ F

D2

cor- respond to R

0

, R

1

∈ R

D1

, R

2

∈ R

D2

, respectively. Suppose, moreover, that x = uv, where u ∈ I(ker D

2

), v ∈ I(ker D

1

) and there are a λ ∈ v

C

R

0

such that Equalities (4.2) hold and a z

0

∈ ker D

1

such that F

1

u = F

1

s

λ

(z

0

) = 0.

Then the initial value problem (4.3)–(4.5) is ill posed and its non-trivial so-

lution is x = uv, where u = s

λ

(z

0

), s

λ

= λR

0

2

I + R

20

), F

1

u = 0 for a

z

0

∈ ker D

1

, v is an eigenvector of R

2

corresponding to the eigenvalue-λ

2

.

(19)

Theorem 4.2. Suppose that X ∈

L

(D

1

)∩

L

(D

2

), Condition [C]

2

is satisfied with respect to the multifunction Ω

(1)1

induced by D

1

, (L

(1)1

, E

1(1)

) ∈ G[Ω

(1)1

], S

(1)

is a sine mapping induced by E

1(1)

, F

0

, F

1

, F

2

are almost averaging initial operators corresponding to R

0

, R

1

∈ R

D1

, R

2

∈ R

D2

, respectively, g

1

= R

0

e, there exists a λ such that iλ ∈ v

C

R

0

, iλg

1

∈ dom (Ω

(1)1

)

−1

, S

(1)

(λg

1

) ∈ ker D

2

and F

1

S

(1)

(λg

1

) = 0. Then the initial value problem (4.3)–(4.5) is ill- posed and its non-trivial solution is x = uv, where u = S

(1)

(λg

1

) ∈ ker D

2

, v ∈ ker D

1

is an eigenvector of R

2

corresponding to the eigenvalue −λ

2

, i.e., 0 6= v ∈ ker(I + λ

2

R

2

).

P roof. Let x = uv. Then, by our assumptions, (D

2

+ λ

2

I)v = D

2

(I + λ

2

R

2

)v = 0. Since both operators D

1

and D

2

satisfy the Leibniz condition and u ∈ ker D

2

, we can apply Corollary 4.1 in a similar way, as before. Since F

0

, F

1

, F

2

are almost averaging and F

0

E

1(j)

(±iλg

1

) = e (cf. [8]), we find

F

0

x = F

0

(uv) = vF

0

u = vF

0

S

(1)

(λg

1

) = v 1 2i F

0

h

E

(1)

(iλg

1

) − E

(1)

(−iλg

1

) i

= 1 2i v h

F

0

E

(1)

(iλg

1

) − F

0

E

(1)

(−iλg

1

) i

= 1

2i v(e − e) = 0;

F

1

x = F

1

(uv) = vF

1

u = vF

1

S

(1)

(λg

1

) = 0;

F

2

x = F

2

(uv) = uF

2

v = uF

2

(−λ

2

R

2

v) = −λ

2

uF

2

R

2

v = 0.

Theorem 4.3. Suppose that X ∈

L

(D

1

)∩

L

(D

2

), Condition [C]

2

is satisfied with respect to the multifunction Ω

(1)1

induced by D

1

, (L

(1)1

, E

1(1)

) ∈ G[Ω

(1)1

], C

(1)

is a cosine mapping induced by E

1(1)

, F

0

, F

1

, F

2

are almost averaging initial operators corresponding to R

0

, R

1

∈ R

D1

, R

2

∈ R

D2

, respectively, g

1

= R

0

e, there exist a λ such that iλ ∈ v

C

R

0

, iλg

1

∈ dom (Ω

(1)1

)

−1

, C

(1)

(λg

1

) ∈ ker D

2

and F

0

C

(1)

(λg

1

) = 0. Then the initial value problem (4.3), (4.4),

(4.6) F

1

Dx = 0

is ill-posed and its non-trivial solution is x = uv, where u = C

(1)

(λg

1

) ∈

ker D

2

, v is an eigenvector of R

2

corresponding to the eigenvalue −λ

2

, i.e.,

0 6= v ∈ ker(I + λ

2

R

2

).

(20)

P roof. Let x = uv. Then, by our assumptions, u ∈ ker D

2

, DC

(1)

(λg

1

) =

−λS

(1)

(λg

1

). Thus, in a similar manner as in the proof of Theorem 4.1, we prove that (D

2

− D

21

)x = 0, F

0

x = 0, F

2

x = 0. Condition F

1

x = 0 follows from the fact that (as before) F

1

is almost averaging, hence F

1

x = vF

1

u = vF

1

C

(1)

(λg

1

)=

1λ

vF

1

DS

(1)

(λ) = 0.

Corollary 4.4. Suppose that all assumptions of Theorem 4.2 are satisfied and F

1

= F

0

, hence also R

1

= R

0

. Then equation (4.3) has a non-trivial solution x = uv, where u = C

(1)

(λg

1

) ∈ ker D

2

, v is an eigenvector of R

2

corresponding to the eigenvalue −λ

2

. This solution satisfies the homogeneous initial conditions

(4.7) F

0

x = 0, F

0

Dx = 0, F

2

x = 0.

Hence the problem (4.3), (4.7) is ill-posed.

Theorem 4.4. Suppose that X ∈

L

(D

1

)∩

L

(D

2

). Condition [C]

2

is satisfied with respect to the multifunction Ω

(2)1

induced by D

2

, (L

(2)1

, E

1(2)

) ∈ G[Ω

(2)1

], S

(2)

, C

(2)

are sine and cosine mappings induced by E

1(2)

, F

0

, F

1

, F

2

are al- most averaging initial operators corresponding to R

0

, R

1

∈ R

D1

, R

2

∈ R

D2

, respectively, g

2

= R

2

e, there exist a λ such that iλ ∈ v

C

R

2

, −λ

2

g

2

∈ dom (Ω

(2)1

)

−1

, z

0

, z

1

∈ ker D

1

, z

2

∈ ker D

2

∩ I(X) and

(4.8) u = z

0

S

(2)

(λg

2

) + z

1

C

(2)

(λg

2

) ∈ I(X).

If v = z

2

E

1(2)

(−λ

2

g

2

) ∈ ker D

1

then x = uv ∈ I(X) is a non-trivial solution of Equation (4.3).

P roof. By the Leibniz condition, v = z

2

E

1(2)

(−λ

2

g

2

) ∈ I(X). Then (D

2

+ λ

2

I)v = D

2

[z

2

E

1(2)

(−λ

2

g

2

) + λ

2

v]

= −λ

2

z

2

E

1(2)

(−λg

2

)D

2

R

2

e + λ

2

v = −λ

2

v + λ

2

v = 0.

By Formulae (3.16), u ∈ ker(D

22

+ λ

2

I). In a similar manner, as in the proof of Corollary 4.1, we get

(D

1

− D

22

)x = (D

1

− D

22

)(uv) = uD

2

v − vD

21

u = u(−λ

2

v) − u(−λ

2

v) = 0.

Cytaty

Powiązane dokumenty

We first iden- tify by means of Design Space Exploration specific GNR topolo- gies for 2- and 3-input {AND, NAND, OR, NOR, XOR, XNOR} and demonstrate by means of the

During each class the topics scheduled below as well as chosen cases and judgements of ECHR (eg. right to life, freedom of expression, right to privacy, right to a fair

If logarithmic mappings exist in X, then cosine and sine elements C(x) and S(x) defined by means of antilogarithmic mappings satisfy the Trigonometric Identity, i.e. also PR[3]):..

They considered continuous self maps of complete metric spaces satisfying one of these contractive conditions and either no additional assumption or one of the following

M´ emoire sur les transcendents elliptiques consid´ er´ ees comme functions de leur amplitudes, Journal d’´ Ecole Royale.

Wydaje się jednak, że stan szkolnictwa parafialnego w diecezji włocławskiej, podobnie jak w wielkopolskiej części diecezji poznańskiej, pogarszał się już w pierwszej połowie

Rice, Lee, Taitt, Boone, Campbell, Walsh, White, Young, Ito Community Development Community Mapping Participatory GIS Community Informatics Telecom Youth & ICTs

Scholen Foto wedstijd Ambassadeurs BespaarProducten Quiz