0% found this document useful (0 votes)
23 views13 pages

10.1515 - Forum 2022 0292

This document summarizes a research article that studies generalized Cauchy-Riemann equations for non-identity bases in finite-dimensional associative commutative algebras with unity. The article provides necessary and sufficient conditions for a set of partial differential equations to characterize the derivative in such an algebra. It generalizes previous work that only considered matrix algebras with canonical basis vectors as the identity. The article proves theorems to characterize all possible matrix algebras that could arise from an algebra's first fundamental representation and relations between partial derivatives when the identity is a linear combination of basis vectors.

Uploaded by

Ali Raza
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
23 views13 pages

10.1515 - Forum 2022 0292

This document summarizes a research article that studies generalized Cauchy-Riemann equations for non-identity bases in finite-dimensional associative commutative algebras with unity. The article provides necessary and sufficient conditions for a set of partial differential equations to characterize the derivative in such an algebra. It generalizes previous work that only considered matrix algebras with canonical basis vectors as the identity. The article proves theorems to characterize all possible matrix algebras that could arise from an algebra's first fundamental representation and relations between partial derivatives when the identity is a linear combination of basis vectors.

Uploaded by

Ali Raza
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 13

Forum Math.

2023; 35(6): 1471–1483

Research Article

Julio Cesar Avila*, Martín Eduardo Frías-Armenta and Elifalet López-González

Generalized Cauchy–Riemann equations in


non-identity bases with application to the
algebrizability of vector fields
https://doi.org/10.1515/forum-2022-0292
Received October 5, 2022; revised January 27, 2023

Abstract: We complete the work done by James A. Ward in the mid-twentieth century on a system of partial
differential equations that defines an algebra 𝔸 for which this system is the generalized Cauchy–Riemann equa-
tions for the derivative introduced by Sheffers at the end of the nineteenth century with respect to 𝔸, which
is also known as the Lorch derivative with respect to 𝔸, and recently simply called 𝔸-differentiability. We get
a characterization of finite-dimensional algebras, which are associative commutative with unity.

Keywords: Generalized Cauchy–Riemann equations, finite-dimensional associative commutative algebras


with unity, Lorch derivative, vector fields

MSC 2020: 35A09, 35F35, 15A27



Communicated by: Jan Frahm

Introduction
The theory of analytic functions in algebras was started by Sheffers [12] at the end of nineteenth century. Other
notable works are [3, 6, 8, 9, 11, 13]. The corresponding differentiability is known as Lorch differentiability which
is associated to algebras 𝔸 (in all this work algebra will be an ℝ-algebra associative commutative with unit), so
we call it 𝔸-differentiability, see Section 2.2. This is similar to how the complex derivative is associated with the
system of complex numbers. We denote by 𝔸 an algebra that as a linear space 𝕃 is ℝn , and by 𝕄 an algebra
that as a linear space 𝕃 is a subspace of matrices of dimension n into M n (ℝ).
In this work, n-dimensional vector field and function from ℝn to ℝn , in both cases differentiable in the
usual sense, have the same meaning, except that we associate integral curves to vector fields. We suppose that
all vector fields are defined on open sets. Although the motivation for the study of algebras comes from the study
of differential equations, in this paper we do not study such differential equations. We will say that a vector field
F is algebrizable if there exists an algebra 𝔸 such that F is 𝔸-differentiable. Next we give a description of part
of our motivation to study the algebrizability of vector fields:
(1) For a vector field and its corresponding system of autonomous ordinary differential equations (ODEs)
f1 ẋ 1 = f1 (x1 , . . . , x n ),
{
{ ..
F = ( ... ) , { . (0.1)
{
fn {ẋ n = f n (x1 , . . . , x n ),

*Corresponding author: Julio Cesar Avila, Escuela de Ingeniería y Ciencias, Tecnologico de Monterrey, Hermosillo, Mexico,
e-mail: jcavila@tec.mx. https://orcid.org/0000-0001-6569-9690
Martín Eduardo Frías-Armenta, Departamento de Matemáticas, Universidad de Sonora, Blvrd. Rosales y Luis Encinas S/N, Col. Centro,
Hermosillo Sonora, C.P. 83000, Mexico, e-mail: eduardo.frias@unison.mx. https://orcid.org/0000-0002-7687-1270
Elifalet López-González, Universidad Autónoma de Ciudad Juárez, Unidad Multidisciplinaria de la UACJ en Cuauhtémoc, Carretera
Cuauhtémoc-Anáhuac, Col. Ejido Anáhuac Km. 3.5 S/N, Mpio. de Cuauhtémoc, Chih., C.P. 31600, Juárez, Mexico,
e-mail: elgonzal@uacj.mx. https://orcid.org/0000-0003-3927-7712

Open Access. © 2023 the author(s), published by De Gruyter. This work is licensed under the Creative Commons Attribution 4.0 Inter-
national License.
1472  J. C. Avila et al., Generalized Cauchy–Riemann equations

we can consider a 𝔸-differential equation of one 𝔸-variable (𝔸-ODE)


dx
= F(x), (0.2)

for which each solution x(τ), which is an 𝔸-differentiable function whose 𝔸-derivative satisfies the differ-
ential equation (0.2), determines a solution ξ(t) = x(te) of the system (0.1). That is, one can use some of the
classical methods to solve ODEs in one variable, to solve equations in one variable over algebras like (0.2)
since one has 𝔸-calculus for 𝔸-differentiability (see [3]), and by evaluating these solutions in the direction
of the identity, te, we obtain solutions of the considered system of ODEs (0.1). Therefore, by solutions of
𝔸-ODEs some systems of ODEs can be solved.
(2) For each 𝔸-algebraizable planar vector field F and each constant b ∈ 𝔸, b ≠ te, where e is the identity
of 𝔸, for all t ∈ ℝ, the vector field G = bF obtained by the product b times F with respect to 𝔸, is a non-
trivial infinitesimal symmetry G of F, so the determinant of the matrix with columns F and G is an inverse
integrating factor of F. See [5] for infinitesimal symmetries and integrating factors. Therefore, an integrating
factor can always be found for algebraizable planar vector fields.
(3) Every algebraizable vector field F is geodesible and the corresponding Riemannian metric tensor g can be
found explicitly. Also, if the vector field is of dimension n, for each regular point of F there exist n − 1 first
integrals whose level sets intersect transversally, whose intersection is a one-dimensional curve which can
be parameterized by arc length with respect to g. Thus the integral curves for these vector fields can be
found. Therefore, ODEs associated with algebraizable vector fields can be solved, see [2] and [4].
(4) For partial differential equations (PDEs) of mathematical physics, families of 𝔸-differentiable functions F
have been found for which there exist linear functions φ such that the families of functions F ∘ φ define
complete solutions of the PDEs, as it is the case of the harmonic functions are related to the conjugate
functions of complex functions, see [7]. Other works related to solutions of PDEs can be seen in [6, 9, 10].
Therefore, 𝔸-differentiable functions give solutions for some PDEs.
(5) For algebrizable vector fields a visualization method for their phase portrait is developed in [1].
For each n-dimensional algebra 𝔸 the 𝔸-differentiability is characterized by a system of n(n − 1) PDEs of first
order, similarly to the complex case, so these systems are called generalized Cauchy–Riemann equations asso-
ciated with 𝔸 (or associated with the 𝔸-differentiability). In the literature on the subject these systems were
assumed to be linearly independent, but no justification for this statement was observed, a proof is given in
Section 5. In [13] an inverse problem arises; given the linearly independent system of PDEs
n n
{ ∑ ∑ d kij f ix j = 0 : 1 ≤ k ≤ n(n − 1)}, (0.3)
j=1 i=1

where d kij represents real constants, f1 , . . . , f n functions of the variables x1 , . . . , x n , and


∂f i
f ix j = ,
∂x j

the question is about the existence of an algebra 𝔸 for which this set is a system of generalized Cauchy–Riemann
equations. In [13], Ward considers matrix algebras 𝕄 that are images 𝕄 = R(𝔸) under the first fundamental
representation R of algebras 𝔸 with unit e = e p in the canonical basis {e1 , e2 , . . . , e n } of ℝn , and solves the
inverse problem for sets (0.3) which are systems of PDEs for these algebras 𝔸. Thus, the general inverse problem
was partially solved. In this paper the work is completed; given a set of PDEs of the type (0.3) we give necessary
and sufficient conditions for the existence of an algebra 𝔸 with unit e = ∑p∈P α p e p , where P ⊂ {1, . . . , n} and
α p ∈ ℝ, such that the given set is a system of Cauchy–Riemann equations for 𝔸 (Theorem 4).
For the proof of Theorem 4 it was necessary to prove a generalization of [13, Ward’s Theorem 1], which
gives sufficient conditions for a set of matrices in M n (ℝ) to be the image of the canonical base of ℝn under
the first fundamental representation of an algebra 𝔸 with unit e = e p in the canonical basis of ℝn . In Section 1
we discuss a condition on how to solve the partial derivatives {f ix j : 1 ≤ i, j ≤ n} in terms of the partial deriva-
tives {f ix p : 1 ≤ i ≤ n} with respect to a single variable x p . The generalization presented in this article, given
in Theorem 1, characterizes all the matrix algebras that are the image of the first fundamental representation
J. C. Avila et al., Generalized Cauchy–Riemann equations  1473

of an algebra 𝔸, and hence their units are not necessarily canonical vectors e i . In this case the solving of the
partial derivatives {f ix j : 1 ≤ i, j ≤ n} of the components is achieved in terms of the partials derivatives of the
components {f ix p : 1 ≤ i ≤ n, p ∈ P} with respect to the variables {x p : p ∈ P} associated with the canonical basis
vectors {e p : p ∈ P} that define the unit
e = ∑ αp ep
p∈P

of 𝔸. Therefore, this characterizes the whole family of algebras.


If all partial derivatives f jx i can be expressed in terms of the partial derivatives f ix p with respect to a single
variable x p , through elementary operations on the system of generalized Cauchy–Riemann equations associ-
ated with an algebra 𝔸, it is possible to arrive at simpler systems of generalized Cauchy–Riemann equations
associated with an algebra 𝔸s , in such a way that the families of functions 𝔸-differentials and 𝔸s -differentials
match. In this way two families of 2D algebras 𝔸 can be constructed. Ward’s work does not consider the set

{f1x2 = 0, f2x1 = 0}, (0.4)

which is a system of generalized Cauchy–Riemann equations for the algebra 𝔸 defined by ℝ2 endowed with
the product between the elements of the canonical basis: e1 e1 = e1 , e1 e2 = 0, e2 e2 = e2 ; 𝔸 has unit e = e1 + e2 .
All other cases of 2D algebras 𝔸 which have unit e = α1 e1 + α2 e2 are already considered in Ward’s work or the
corresponding Cauchy–Riemann equations are equivalent to a system already considered by Ward’s work, see
Section 6 and [4]. Example 4 illustrates this for the case of 3D algebras. If we add this algebra that is missing in
Ward’s work, we obtain three families of two-dimensional algebras such that each algebrizable vector field is
𝔸-differentiable for an algebra 𝔸 in some of these families. This has been useful in the following two contexts:
in the study of vector fields which are differentiable in the sense of Lorch, see [4], and in the construction of
complete solutions of families of PDEs of the type

∂2 u
Au x1 x1 + Bu x1 x2 + Cu x2 x2 = 0, u xi xj = ,
∂x i x j

which generalizes the classical result showing that the components of complex analytic functions define a com-
plete solution of the 2D Laplace equation, see [7]. There are other papers that have worked on the solution of
PDEs of mathematical physics through algebras, see [9], [10], and references therein.
In Theorem 2 we present three equivalences of 𝔸-differentiability: item (2) is the generalization of classic
Cauchy–Riemann equations F2 = iF1 , item (3) was presented by Sheffers in [12, Satz 3], in form of components,
item (4) is a generalization of [13, equation (18), p. 460]. For the algebras characterized in Theorem 1 the gener-
alized Cauchy–Riemann equations given in Theorem 2 give a characterization of the algebrizability of vector
fields. That is, a vector field F is algebrizable if and only if there exists an algebra 𝔸, which is given in Theorem 1,
whose associated Cauchy–Riemann equations, given in Theorem 2, are satisfied by F.
All the results obtained in this paper are made over the real field ℝ, however they can be generalized to
any field 𝔽, as it is made in Ward’s paper [13].

1 Ward’s paper
Definition 1. We will say that system (0.3) satisfies the zero trace condition if ∑ni=1 d kii = 0.

In Ward’s work [13], systems of n(n − 1) first-order linear PDEs of the form (0.3) are considered. Ward’s approach
is about the existence of an algebra 𝔸 such that the set of equations (0.3) is a system of generalized Cauchy–
Riemann equations for the 𝔸-derivative. One of the conditions required by Ward is the existence of a variable
x p ∈ {x1 , x2 , . . . , x n } such that all the partial derivatives f ix j , for 1 ≤ i, j ≤ n, can be solved in terms of a linear
combination of partial derivatives of the set {f ix p : i = 1, 2, . . . , n}. From system (0.3) for each x j there is a matrix
M j ∈ M n(n−1),n (ℝ) such that system (0.3) is written as

M1 F x1 + ⋅ ⋅ ⋅ + M n F x n = 0. (1.1)
1474  J. C. Avila et al., Generalized Cauchy–Riemann equations

Ward’s condition above implies the existence of matrices A i ∈ M n (ℝ) such that
f1x1 ⋅⋅⋅ f1x n
. .. ..
( .. . . ) = f1x p A1 + ⋅ ⋅ ⋅ + f nx p A n . (1.2)
f nx1 ⋅⋅⋅ f nx n
A second condition is the commutativity of the set {A1 , . . . , A n }, and a third condition is that system (0.3) satisfies
the zero trace condition. Under these three conditions Ward [13] proved the existence of the algebra 𝔸 with the
required conditions.
The solution condition of all the partial derivatives f ix j of the components f i in terms of the partial deriva-
tives f ix p of the components f i with respect to a single variable x p , reduces to verifying the invertibility of n
matrices of n(n − 1) × n(n − 1), as we see below.
Consider the matrix M ∈ M n(n−1),n2 (ℝ) given by
M = (M1 M2 ⋅⋅⋅ Mn ) . (1.3)
Denote by π i : M n(n−1),n2 (ℝ) → M n(n−1),n(n−1) (ℝ) the projection which avoid the i-th sumbatrix M i from M
π i (M) = (M1 M2 ⋅⋅⋅ M i−1 M i+1 ⋅⋅⋅ Mn ) .
The following proposition gives conditions under which there exists p such that equality (1.2) is satisfied.

Proposition 1.1. If for some p the matrix π p (M), where M is given in (1.3), is invertible, then all partial derivatives
in {f ix j : 1 ≤ i, j ≤ n} can be written in terms of partial derivatives in {f ix p : 1 ≤ i ≤ n}.

Proof. One can start from a system as (1.1) and if some matrix π p (M) is invertible, then multiplying the system
by π p (M)−1 we get a new system. Thus, the partial derivatives can be written by
F x1
.. f1x i
. f2x i
( )
(F x p−1 ) −1 ( .. )
( )
(F ) = −π p (M) M p F p , Fi = ( . ) . (1.4)
( x p+1 )
.. f n−1x i
.
( f nx i )
( F xn )
Thus, the proof is finished.

2 Algebras and 𝔸-differentiability


2.1 Algebras and matrix algebras
Definition 2. We call an ℝ-linear space 𝕃 an algebra if it is endowed with a bilinear product 𝕃 × 𝕃 → 𝕃 denoted
by (x, y) 󳨃→ xy, which is associative and commutative x(yz) = (xy)z and xy = yx for all x, y, z ∈ 𝕃; furthermore,
there exists a unit e ∈ 𝕃, which satisfies ex = x for all x ∈ 𝕃.

An algebra 𝕃 will be denoted by 𝔸 if 𝕃 = ℝn and by 𝕄 if 𝕃 is an n-dimensional matrix algebra in the space of


matrices M(n, ℝ), where the algebra product corresponds to the matrix product.

Definition 3. If 𝔸 is an algebra, the 𝔸-product between the elements of the canonical basis {e1 , e2 , . . . , e n } of
ℝn is given by
n
e i e j = ∑ c ijk e k ,
k=1
where c ijk ∈ ℝ for i, j, k ∈ {1, 2, . . . , n} are called structure constants of 𝔸. The first fundamental representation
of 𝔸 is the injective linear homomorphism R : 𝔸 → M(n, ℝ) defined by R : e i 󳨃→ R i , where R i is the matrix with
[R i ]jk = c ikj , for i = 1, 2, . . . , n.
J. C. Avila et al., Generalized Cauchy–Riemann equations  1475

2.2 𝔸-differentiability and algebrizability of vector fields

The 𝔸-differentiability of vector fields is the same definition as the differentiability in the sense of Lorch with
respect to 𝔸, see [8].

Definition 4. Let 𝔸 be an algebra, and F a vector field which is defined and differentiable in the usual sense on
an open set Ω ⊂ ℝn . We say F is 𝔸-differentiable on Ω if there exists a vector field F 󸀠 defined on Ω such that

dF p (v) = F 󸀠 (p) ⋅ v, (2.1)

where F 󸀠 (p) ⋅ v denotes the 𝔸-product of F 󸀠 (p) and v for every vector v in ℝn and p ∈ Ω.

For the 𝔸-differentiability, most of the known results on calculus in ℝ or ℂ transfers to 𝔸-calculus, see [3], only
one must to be careful with singular elements, these are non-invertible elements with respect to the 𝔸-product.

Definition 5. We say two system of linear partial differential equation (PDEs) with constant coefficients are
equivalent if through elementary row operations carry one of them to the other.

The 𝔸-differentiability has associated sets of PDEs, see Theorem 2.

Definition 6. We call generalized Cauchy–Riemann equations associated to 𝔸 to any system of PDEs equivalent
to equations obtained of e j F i = e i F j , with i, j ⊂ {1, 2, . . . , n}.

3 Characterization of algebras
The following theorem, proved in [13] for P with |P| = 1 and α p = 1, characterizes the associative commuta-
tive algebras 𝔸 with unit e p in the canonical basis {e1 , e2 , . . . , e n } of ℝn . This completes the characterization
of associative commutative algebras 𝕄 in M n (ℝ) that are the image of a first fundamental representation of
n-dimensional algebras 𝔸. Since algebras are isomorphic to their first fundamental representations, this gives
a complete characterization of the algebras. Also this result is used to give conditions on PDEs systems so that
they are generalized Cauchy–Riemann equations.

Theorem 1. The spanned set by {A i : i = 1, . . . , n} is the image of the first fundamental representation of an
algebra 𝔸, with R(e i ) = A i , e = ∑p∈P α p e p , where
n
A i A j = ∑ a ijt A t , I = ∑ αp Ap , (3.1)
t=1 p∈P

if and only if
(a) there exists a commutative set {A1 , A2 , . . . , A n } ⊂ M n (ℝ), where A i = (a isr ), that is

Ai Aj = Aj Ai , i, j = 1, . . . , n, (3.2)

(b) there exists an index set P ⊂ {1, . . . , n} with {α p }p∈P such that

∑ α p a ipr = δ ir , i, r = 1, . . . , n. (3.3)
p∈P

Proof. The proof in the forward direction is known, see [6, p. 642, equation 4].
Conversely, let B ij = A i A j and let b pu the element of the matrix B ij with row-index u and column-index p.
Then by (3.3)
∑ α p b pu = ∑ α p ∑ a itu a jpt = ∑ a itu ∑ α p a jpt = ∑ a itu δ jt = a iju .
p∈P p∈P t=1 t=1 p∈P t=1

Using that A i A j = A j A i , and doing the same calculations as above, for A j A i we have

a iju = a jiu for i, j, u = 1, . . . , n. (3.4)


1476  J. C. Avila et al., Generalized Cauchy–Riemann equations

Furthermore, another expression for the entries of A i A j = A j A i is

∑ a itu a jvt = ∑ a jtu a ivt . (3.5)


t=1 t=1

Then we have (following Ward’s proof [13])


n n n n n n
(A i A j )rs = ∑ a itr a jst = ∑ a jtr a ist = ∑ a jtr a sit = ∑ a str a jit = ∑ a tsr a ijt = ∑ a ijt a tsr
t=1 t=1 t=1 t=1 t=1 t=1

and then A i A j =∑nt=1 a ijt A t . The first equality is obtained by matrix-product definition, the second and fourth
are by (3.5), the third and fifth by (3.4), and the sixth by commutativity of ℝ. From (3.3), we see that the A i are
linearly independent with respect to ℝ. Now we shall prove ∑p∈P α p A p = I. If
a p11 a p21 ⋅⋅⋅ a pn1
a p12 a p22 ⋅⋅⋅ a pn2
Ap = ( . .. .. .. ) ,
.. . . .
a p1n a p2n ⋅⋅⋅ a pnn
then analyzing every element of the matrix ∑p∈P α p A p (with row-index r and column-index s)

( ∑ αp Ap ) = ∑ α p a psr = ∑ α p a spr = δ sr ,
p∈P rs p∈P p∈P

and then ∑p∈P α p A p = I, where we used (3.4) and (3.3).


Next, an example outside the scope of Theorem 1 is given, i.e., the algebra is not image of a first fundamental
representation.

Example 1. Consider the matrices β = {A1 , A2 , A3 } given by


1
2 0 0 0 0 0 0 0 0
1
A1 = ( 0 3 0) , A2 = (0 − 13 0 ), A 3 = (0 0 0) .
1
0 0 3 0 0 − 31 0 1 0
It can be verified that the matrices {A1 , A2 , A3 } are commutative, and their matrix products satisfy the following
relations:
A1 A2 A3
1 1 1 1
A1 2 A1 + 6 A2 3 A2 3 A3
1
. (3.6)
A2 3 A2 − 31 A2 − 31 A3
1
A3 3 A3 − 31 A3 0
Then they define a 3D commutative matrix algebra 𝕄, which in this case is given by

{ x 0 0 }
{ }
𝕄 = {(0 y 0) : x, y, z ∈ ℝ} .
{ }
{ 0 z y }
We take P = {1, 2}, α1 = 2, and α2 = −1, since 2a112 − a122 = 2 ⋅ 0 − 31 = − 13 , the conditions of Theorem 1 are not
satisfied. Then it is not first fundamental representation.
We can find the first fundamental representation with respect to this basis, which would give a matrix
algebra 𝕄R , which should not match 𝕄, but should be an algebra of simultaneously diagonalizable matrices
which is conjugate to 𝕄, i.e., 𝕄 = B𝕄R B−1 , where B is an invertible matrix.

Next, two examples are given where the algebra is the image of a first fundamental representation.

Example 2. The following matrices satisfy (3.2) and (3.3) of Theorem 1


1
2 0 0 0 0 0 0 0 0
1 1
A1 = (6 3 0) , A2 = ( 31 − 13 0 ), A3 = ( 0 0 0) .
1
0 0 3 0 0 − 31 1
3 − 13 0
J. C. Avila et al., Generalized Cauchy–Riemann equations  1477

Actually, they have the same matrix products as (3.6). For condition (3.3) we take P = {1, 2}, with α1 = 2 and
α2 = −1. Then
2a111 − a121 2a211 − a221 2a311 − a321 1 0 0
2A1 − A2 = (2a112 − a122 2a212 − a222 2a312 − a322 ) = (0 1 0) .
2a113 − a123 2a213 − a223 2a313 − a323 0 0 1
Example 3. Consider the matrices β = {A1 , A2 , A3 } given by

1 0 0 0 0 0 0 0 0
A1 = (0 0 0) , A2 = (0 1 0) , A 3 = (0 0 0) .
0 0 0 0 0 1 0 1 0
We take P = {1, 2}, α1 = 1, and α2 = 1. Then
a111 + a121 a211 + a221 a311 + a321 1 0 0
(a112 + a122 a212 + a222 a312 + a322 ) = (0 1 0) .
a113 + a123 a213 + a223 a313 + a323 0 0 1
Therefore, the conditions of Theorem 1 are satisfied.
It can be verified that the matrices {A1 , A2 , A3 } are commutative, I = A1 + A2 with I the identity matrix, and
their matrix products satisfy the following relations:

A1 A2 A3
A1 A1 0 0
.
A2 0 A2 A3
A3 0 A3 0
Thus, R(A i ) = A i for i = 1, 2, 3.

4 Characterization of algebrizable vector fields


In the following lemma we think the elements of ℝn as columns.

Lemma 4.1. Let 𝔸 be an algebra and R : 𝔸 → M n (ℝ) its first fundamental representation. Then R(a)b = ab,
where R(a)b denotes the product between the matrix R(a) and the vector b, and ab denotes the product in 𝔸.

Proof. Firstly, we see that R(e i )e j = e i e j :

c i11 c i21 ⋅⋅⋅ c in1 c ij1


c i12 c i22 ⋅⋅⋅ c in2 c ij2 n
R(e i )e j = ( . .. ) e = ( ) = ∑ c ijk e k .
.. .
..
. ... j ..
. k=1
c i1n c i2n ⋅⋅⋅ c inn c ijn
Then
n n n n
R(e i )b = R(e i ) ∑ b j e j = ∑ b j R(e i )e j = ∑ b j e i e j = e i ∑ b j e j = e i b.
j=1 j=1 j=1 j=1

Next, using the previous equality we obtain


n n n
R(a)b = R( ∑ a i e i )b = ∑ a i R(e i )b = ∑ a i e i b = ab.
i=1 i=1 i=1

This prove the lemma.


By using Lemma 4.1, the Cauchy–Riemann equations e i F k = e k F i can be written as

R(e i )F k = R(e k )F i , i, i ∈ {1, 2, . . . , n}, i ≠ j.


1478  J. C. Avila et al., Generalized Cauchy–Riemann equations

If the unit e of 𝔸 is given by e = ∑p∈P α p e p , then the partial derivatives f ix k of the components f i of algebraiz-
able vector fields F can be expressed as a linear combination of {f ix p : p ∈ P}, this is included in the fourth
characterization of 𝔸-differentiability given in the following theorem. Note that (3) is the same as [12, Satz 3].

Theorem 2. Let F = (f1 , f2 , . . . , f n ) be a differentiable field vector in the usual sense, 𝔸 an algebra with first fun-
damental representation R given by R(e i ) = R i and unity e = ∑p∈P α p e p , where P is an index set P ⊂ {1, . . . , n}.
Then the following items are equivalent:
(1) F is 𝔸-differentiable.
(2) F satisfies e j F x i = e i F x j for all i, j ∈ {1, 2, . . . , n} with respect to 𝔸.
(3) The partial derivatives F x k of F satisfy
F xk = R k ∑ α p F xp . (4.1)
p∈P

(4) The Jacobian of F satisfies

JF = ∑ α p f1x p R1 + ∑ α p f2x p R2 + ⋅ ⋅ ⋅ + ∑ α p f nx p R n . (4.2)


p∈P p∈P p∈P

Proof. (1) ⇒ (2) Since F is 𝔸-differentiable, we have that there exists a vector field F 󸀠 such that dF x (v) = F 󸀠 (x)v
for every vector v. This implies that

e j F x i = e j dF(e i ) = e j F 󸀠 e i = e i F 󸀠 e j = e i dF(e j ) = e i F x j ,

which are the generalized Cauchy–Riemann equations associated to 𝔸.


(2) ⇒ (3) If
R i F x k = R k F x i for i, k = 1, . . . , n,
then
α p R p F xk = α p R k F xp for p ∈ P.
Thus, summing for p ∈ P,
∑ α p R p F xk = ∑ α p R k F xp ,
p∈P p∈P

F xk = R k ∑ α p F xp ,
p∈P

where ∑p∈P α p R p = I is the identity matrix because the expression of the identity e.
(3) ⇒ (4) Let U = ∑p∈P α p F x p be. From (3) we have F x k = R k U. Thus, the Jacobian matrix JF of F is given in
component notation by

JF = ( R1 U R2 U ... Rn U )
r111 r121 ⋅⋅⋅ r1n1 u1 r n11 r n21 ⋅⋅⋅ r nn1 u1
r112 r122 ⋅⋅⋅ r1n2 u2 r n12 r n22 ⋅⋅⋅ r nn2 u2
= (( . .. .. .. ) ( .. ) ⋅⋅⋅ ( . .. .. .. ) ( .. ))
.. . . . . .. . . . .
r11n r12n ⋅⋅⋅ r1nn un r n1n r n2n ⋅⋅⋅ r nnn un
r1i1 r2i1 r ni1
n r1i2 n r2i2 n r ni2
= (∑ u i ( . ) ∑ ui ( . ) ⋅⋅⋅ ∑ u i ( . ))
.. .. ..
i=1 i=1 i=1
r1in r2in r nin
r1i1 r2i1 ⋅⋅⋅ r ni1 r i11 r i21 ⋅⋅⋅ r in1
n r1i2 r2i2 ⋅⋅⋅ r ni2 n r i12 r i22 ⋅⋅⋅ r in2
= ∑ ui ( . .. .. .. ) = ∑ u i ( .. .. .. .. ) ,
.. . . . . . . .
i=1 i=1
r1in r2in ⋅⋅⋅ r nin r i1n r i2n ⋅⋅⋅ r inn

where the last equality is obtained from commutativity of 𝔸.


J. C. Avila et al., Generalized Cauchy–Riemann equations  1479

(4) ⇒ (1) Suppose that the Jacobian of F satisfies equality (4.2). So our candidate for 𝔸-derivative of F is
n
∑ ∑ α p f kx p e k .
k=1 p∈P

We have to prove the equality


n
dF(v) = ( ∑ ∑ α p f kx p e k )(v), (4.3)
k=1 p∈P

where the product indicated on the right hand side represents the product of 𝔸. First, we have that the differ-
ential of F applied to v is the Jacobian matrix JF of F multiplied by v through the matrix product
n
dF(v) = ( ∑ ( ∑ α p f kx p )R k )(v). (4.4)
k=1 p∈P

Next, by Lemma 4.1 we have


n
dF(v) = ∑ ( ∑ α p f kx p )(e k v), (4.5)
k=1 p∈P

where e k v represent the product with respect to an 𝔸. Therefore equality (4.3) holds, because the right side
of (4.5) is equal to the right side of (4.3). This shows that F is 𝔸-differentiable and that the 𝔸-derivative of F is
n
F 󸀠 = ∑ ( ∑ α p f kx p )e k .
k=1 p∈P

Due to Theorem 2, in the next corollary we give the set of all solutions of system (0.3).

Corollary 4.1. If there is an algebra 𝔸 such that system (0.3) are the Cauchy–Riemann equations for 𝔸, then the
𝔸-differentiable functions are all solutions of system (0.3).

The following example gives two algebras 𝔸1 and 𝔸2 for which the family of functions 𝔸1 -differentiable and
𝔸2 -differentiable are the same, since they have the same generalized Cauchy–Riemann equations.

Example 4. The linear space ℝ3 endowed with the product


e1 e2 e3
1 1 1 1
e1 2 e1 + 6 e2 3 e2 3 e3
1
,
e2 3 e2 − 31 e2 − 31 e3
1
e3 3 e3 − 31 e3 0
define an algebra 𝔸 with unit e = 2e1 − e2 . The Cauchy–Riemann equations for the 𝔸-derivative are given by
f1y = 0, f1x − f2x − f2y = 0, f3x + f3y = 0,
f1z = 0, f2z = 0, f1x − f2x − f3z = 0.
Thus,
1
f1x f1y f1z 2 0 0 0 0 0 0 0 0
1 1
(f2x f2y f2z ) = (2f1x − f1y ) ( 6 3 0 ) + (2f2x − f2y ) ( 31 − 13 0 ) + (2f3x − f3y ) ( 0 0 0) .
1
f3x f3y f3z 0 0 3 0 0 − 31 1
3 − 13 0
That is,
JF = (2f1x − f1y )R1 + (2f2x − f2y )R2 + (2f3x − f3y )R3 ,
where R i = R(e i ), R : 𝔸 → M n (ℝ) of 𝔸.
On the other hand, for the same generalized Cauchy–Riemann equations, the partial derivatives f ix j can be
written in terms of a linear combination of f1x , f2x , f3x , from which we obtain
f1x f1y f1z 1 0 0 0 0 0 0 0 0
(f2x f2y f2z ) = f1x (0 1 0) + f2x (1 −1 0 ) + f3x (0 0 0) .
f3x f3y f3z 0 0 1 0 0 −1 1 −1 0
1480  J. C. Avila et al., Generalized Cauchy–Riemann equations

Since these matrices satisfy the Ward conditions, we have that F = (f1 , f2 , f3 ) is 𝔹-differentiable, where 𝔹 is the
algebra defined by ℝ3 with respect to the product
e1 e2 e3
e1 e1 e2 e3
.
e2 e2 −e2 −e3
e3 e3 −e3 0
The unit e of 𝔹 is e = e1 .

5 Linear independence of Cauchy–Riemann equations


Theorem 3. A set of generalized Cauchy–Riemann equations associated with an algebra 𝔸 contains n(n − 1)
linearly independent PDEs.

Proof. By Theorem 2, a set of generalized Cauchy–Riemann equations associated with an algebra 𝔸 is equivalent
to equation (4.1). Then a set of Cauchy–Riemann equations for 𝔸-differentiability is given by
n
Fk = Rk ∑ αp Fp , k = 1, 2, . . . , n,
p=1

where R i are its first fundamental representation n × n matrices with

I = α1 R1 + ⋅ ⋅ ⋅ + α n R n .

They can be rewritten as n equations as follows:


F1 = R1 (α1 F1 + ⋅ ⋅ ⋅ + α n F n ),
..
.
F n = R n (α1 F1 + ⋅ ⋅ ⋅ + α n F n ),
as well as
(α1 R1 − I)F1 + α2 R1 F2 + ⋅ ⋅ ⋅ + α n R1 F n = 0,
..
.
α1 R n F1 + α2 R n F2 + ⋅ ⋅ ⋅ + (α n R n − I)F n = 0.
In order to prove that the later equations are linearly independent, we are to consider the next n × n-matrix
(where each entry is another n × n-matrix)
α1 R1 − I α2 R1 α3 R1 ⋅⋅⋅ α n R1
[ α R α2 R2 − I α3 R2 ⋅⋅⋅ α n R2 ]
[ 1 2 ]
[ .. .. .. .. .. ],
[ . ]
[ . . . . ]
[ 1 Rn
α α2 R n α3 R n ⋅⋅⋅ αn Rn − I]
and prove that this matrix has maximal range, namely n(n − 1). For this it is only necessary to prove that only
one of the columns is linearly dependent of the others. To achieve this, we will do operations between rows and
columns in order to preserve the same set of solutions, and one column will be only zeros.
For this, we take only the columns where α p ≠ 0, because if α p = 0, then that column will have only zeros
except the p-th entry, as follows:
0
[.]
[.]
[.]
[ ]
[−I ] .
[ ]
[.]
[.]
[.]
[0]
J. C. Avila et al., Generalized Cauchy–Riemann equations  1481

Therefore, let us consider the non-zero α p , let us say α p for p = 1, 2, . . . , l. Then the n × l-matrix with α p ≠ 0 is
α1 R1 − I α2 R1 α3 R1 ⋅⋅⋅ α l R1
[ α R α2 R2 − I α3 R2 ⋅⋅⋅ α l R2 ]
[ 1 2 ]
[ .. .. .. .. .. ].
[ . ]
[ . . . . ]
[ α1 R l α2 R l α3 R l ⋅⋅⋅ αl Rl − I]
To the later matrix, it will be done the next row operation: fixing the first row, for each k-th row, k = 2, 3, . . . , l,
k-th row is replaced by k-th row plus ααk1 times first row. Then one gets
α1 R1 − I α2 R1 α3 R1 ⋅⋅⋅ α l R1
[ α1 α3 αl ]
[ α2 (α1 R1 + α2 R2 − I) α1 R1 + α2 R2 − I α2 (α 1 R 1 + α 2 R 2 ) ⋅⋅⋅ α2 (α 1 R 1 + α 2 R 2 )]
[ ]
[ α1 (α R + α R − I) α2 αl ]
[ α3 1 1
[
3 3 α3 (α 1 R 1 + α 3 R 3 ) α1 R1 + α3 R3 − I ⋅⋅⋅ α3 (α 1 R 1 + α 3 R 3 )] .
]
[ .. .. .. .. .. ]
[ . . . . . ]
α1 α2 α3
[ α l (α 1 R 1 + α l R l − I) α l (α 1 R 1 + αl Rl ) α l (α 1 R 1 + αl Rl ) ⋅⋅⋅ α1 R1 + α l R l − I ]
Now, for the next column operation: fix the last column, for each s-th column, s = 1, 2, . . . , l − 1, the s-th column
is replaced by s-th column minus ααsl times last column. Then one gets
−I 0 0 ⋅⋅⋅ α l R1
[ α1 αl ]
[− α 2 I −I 0 ⋅⋅⋅ α2 (α 1 1 + α 2 R 2 )]
R
[ ]
[− α1 I 0 αl ]
[ α3
[ .
−I ⋅⋅⋅ α3 (α 1 R 1 + α 3 R 3 )] . ]
[ . .. .. .. .. ]
[ . . . . . ]
α2 α3
[ 0 αl I αl I ⋅⋅⋅ α1 R1 + α l R l − I ]
This last matrix is almost an inferior triangular matrix, except for the last column. Now, we are going to fix
the first row, and for each k-th row, k = 2, 3, . . . , l, k-th row is replaced by k-th row minus ααk1 times first row.
Then one gets
−I 0 0 ⋅⋅⋅ α l R1
[ 0 −I 0 ⋅ ⋅ ⋅ α l R2 ]
[ ]
[ ]
[ 0 0 −I ⋅ ⋅ ⋅ α l R3 ] .
[ . .. .. .. ]
[ . .. ]
[ . . . . . ]
α1 α2 α3
[ αl I αl I αl I ⋅ ⋅ ⋅ α R
l l − I ]
Finally, we are going to do several row operations at once, we are going to sum a multiple constant of each row,
to the last row, i.e. the last row will be replaced by
α1 α2 α l−1
l-th row + (first row) + (second row) + ⋅ ⋅ ⋅ + [(l − 1) − th row].
αl αl αl
Then one gets
−I 0 0 ⋅⋅⋅ α l R1
[0 −I 0 ⋅⋅⋅ α l R2 ]
[ ]
[ ]
[0 0 −I ⋅⋅⋅ α l R3 ] .
[. .. .. .. ]
[. .. ]
[. . . . . ]
[0 0 ] 0 0 ⋅⋅⋅
Where we use the expression I = α1 R l + α2 R2 + ⋅ ⋅ ⋅ + α l R l in the (l, l)-entry. Therefore, the range of the last
matrix is (l − 1) and each entry is a n × n matrix. In addition, we have (n − l) independent columns because the
α p = 0. With this we have shown that the n(n − 1) Cauchy–Riemann equations are linearly independent.

6 Ward completion
The next theorem is the main result in Ward’s paper [13].
1482  J. C. Avila et al., Generalized Cauchy–Riemann equations

Theorem (Ward [13, Theorem 2]). Suppose the system of PDEs (0.3) has the property that for some fixed integer p,
it implies the set
JF = f1x p R1 + f2x p R2 + ⋅ ⋅ ⋅ + f nx p R n .
Suppose further that the matrices A i = (a isr ), i = 1, . . . , n, are commutative and satisfy the zero trace condition.
Then there is a uniquely determined algebra 𝔸 for which (0.3) is a set of generalized Cauchy–Riemann differential
equations.

Now, we give a generalization of the algebrizability of vector fields, in the following theorem, which completes
the above theorem.

Theorem 4. There exists an algebra 𝔸 for which the set (0.3) is the system of generalized Cauchy–Riemann equa-
tions if and only if the following three statements are satisfied:
(1) there exists a set of matrices {A i = (a isr ) : i = 1, 2, . . . , n} in M n (ℝ) such that set (0.3) implies equality (4.2)
for R i = (a isr ),
(2) {A i : i = 1, 2, . . . , n} is commutative, that is, A i A j = A j A i for 1 ≤ i, j ≤ n, and
(3) set (0.3) satisfies the zero trace condition (Definition 1).

Proof. Suppose we take 𝔸 such that the set of PDEs (0.3) is its system of generalized Cauchy–Riemann equations,
and A i = R(e i ) for i = 1, 2 . . . , n, where R is the first fundamental representation of 𝔸. Thus:
(1) is a direct consequence of Theorem 3,
(2) is satisfied because is 𝔸 is an algebra,
(3) system (0.3) satisfies the zero trace condition because it is satisfied by the set of PDEs obtained by equality
(4.2) of Theorem 2 and this set is equivalent to system (0.3).
Now we have to show the converse.
Since each PDE of system (4.2) is a linear combination of the set of PDEs (0.3), it follows since PDEs (0.3) hold
the zero trace condition of PDEs (0.3), that is, ∑ni=1 d kii = 0 with k = 1, 2, . . . , n(n − 1), thus ∑p∈P α p a tpr = δ rt .
For example, (4.2) can be rewritten as

f11 f12 ... f1n a111 a121 ... a1n1


f21 f22 ... f2n a112 a122 ... a1n2
0 = −( . .. .. .. ) + ( .. .. .. .. ) ∑ α p f1p
.. . . . . . . . p∈P
f n1 f n2 ... f nn a11n a12n ... a1nn
a211 a221 ... a2n1 a n11 a n21 ... a nn1
a212 a222 ... a2n2 a n12 a n22 ... a nn2
+( .
.. .. .. .. ) ∑ α p f2p + ⋅ ⋅ ⋅ + ( .. .. .. .. ) ∑ α p f np .
. . . p∈P . . . . p∈P
a21n a22n ... a2nn a n1n a n2n ... a nnn

Then, for k = 1 one has

0 = −f11 + a111 (α1 f11 + ⋅ ⋅ ⋅ + α n f n ) + a211 (α1 f11 + ⋅ ⋅ ⋅ + α n f n ) + ⋅ ⋅ ⋅ + a n11 (α1 f11 + ⋅ ⋅ ⋅ + α n f n )

and the coefficients are d111 = −1 + α1 a111 , d122 = α2 a211 , . . . , d1nn = α n a n11 . Then from the zero trace condition
of PDEs (0.3),
n n n
∑ d1ii = −1 + ∑ α i a i11 = −1 + ∑ α i a1i1 = 0. (6.1)
i=1 i=1 i=1

Similarly, for k = 2 one has

0 = −f12 + a121 (α1 f11 + ⋅ ⋅ ⋅ + α n f n ) + a221 (α1 f11 + ⋅ ⋅ ⋅ + α n f n ) + ⋅ ⋅ ⋅ + a n21 (α1 f11 + ⋅ ⋅ ⋅ + α n f n )

and the coefficients are d211 = α1 a121 , d222 = α2 a221 , . . . , d2nn = α n a n21 . Then from the zero trace condition of
PDEs (0.3),
n n n
∑ d kii = ∑ α i a i21 = ∑ α i a2i1 = 0. (6.2)
i=1 i=1 i=1
J. C. Avila et al., Generalized Cauchy–Riemann equations  1483

In both cases (6.1)–(6.2), we used commutativity and ∑p∈P α p a tpr = δ rt . Hence by Theorem 1 the A i , i = 1, 2, . . . , n,
form a basis of an algebra 𝕄 with identity I = ∑p∈P α p A p . Therefore, if 𝔸 is the first fundamental representation
of 𝕄, the set of PDEs (0.3) is satisfied for the 𝔸-differentiable functions.

The following corollary is given by Ward in [13].

Corollary 6.1. A necessary and sufficient condition that the linearly independent PDEs
2
∑ d kij f ix j = 0, k = 1, 2, (6.3)
i,j=1

determine an algebra 𝔸 for which (6.3) is a set of generalized Cauchy–Riemann equations is that

d k11 + d k22 = 0, k = 1, 2. (6.4)

A system of generalized Cauchy–Riemann equations for the algebra 𝔸 with product given in the canonical basis
of ℝ2 given by e1 e1 = e1 , e1 e2 = 0, e2 e2 = e2 , is the set (0.4). The unit e of 𝔸 is e = e1 + e2 . The 𝔸-differentiable
functions is the set of all the functions F = (f1 , f2 ), where f1 (x1 , x2 ) = f(x1 ), f2 (x1 , x2 ) = g(x2 ), and f , g are differ-
entiable functions of one variable. Therefore, this is a case not covered by [13, Theorem 2]. However, this system
satisfies and is in harmony with Corollary 6.1.

References
[1] A. Alvarez-Parrilla, M. E. Frías-Armenta, E. López-González and C. Yee-Romero, On solving systems of autonomous ordinary
differential equations by reduction to a variable of an algebra, Int. J. Math. Math. Sci. 2012 (2012), Article ID 753916.
[2] J. C. Avila, M. E. Frías-Armenta and E. López-González, Geodesibility of algebrizable three-dimensional vector fields, in preparation.
[3] J. Cook, Introduction to A-calculus, preprint (2017), https://arxiv.org/abs/1708.04135.
[4] M. E. Frías-Armenta and E. López-González, On geodesibility of algebrizable planar vector fields, Bol. Soc. Mat. Mex. (3) 25 (2019),
no. 1, 163–186.
[5] I. A. García and M. Grau, A survey on the inverse integrating factor, Qual. Theory Dyn. Syst. 9 (2010), no. 1–2, 115–166.
[6] P. W. Ketchum, Analytic functions of hypercomplex variables, Trans. Amer. Math. Soc. 30 (1928), no. 4, 641–667.
[7] E. López-González, On solutions of PDEs by using algebras, Math. Methods Appl. Sci. 45 (2022), no. 8, 4834–4852.
[8] E. R. Lorch, The theory of analytic functions in normed Abelian vector rings, Trans. Amer. Math. Soc. 54 (1943), 414–425.
[9] S. A. Plaksa, Monogenic functions in commutative algebras associated with classical equations of mathematical physics, Ukr. Mat.
Visn. 15 (2018), no. 4, 543–575.
[10] A. Pogorui, R. M. Rodríguez-Dagnino and M. Shapiro, Solutions for PDEs with constant coefficients and derivability of functions
ranged in commutative algebras, Math. Methods Appl. Sci. 37 (2014), no. 17, 2799–2810.
[11] S. V. Rogosin and A. A. Koroleva, Advances in Applied Analysis, Trends Math., Birkhäuser/Springer, Basel, 2012.
[12] G. Scheffers, Verallgemeinerung der Grundlagen der gewöhnlich komplexen Funktionen I, Leipz. Ber. 45 (1893), 828–848.
[13] J. A. Ward, From generalized Cauchy–Riemann equations to linear algebras, Proc. Amer. Math. Soc. 4 (1953), 456–461.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy