Bicomplex Holomorphic Functions - The Algebra, Geometry and Analysis of Bicomplex Numbers (PDFDrive)
Bicomplex Holomorphic Functions - The Algebra, Geometry and Analysis of Bicomplex Numbers (PDFDrive)
M. Elena Luna-Elizarrarás
Michael Shapiro
Daniele C. Struppa
Adrian Vajiac
Bicomplex
Holomorphic Functions:
The Algebra,
Geometry and Analysis
of Bicomplex
Numbers
Frontiers in Mathematics
Bicomplex Holomorphic
Functions
The Algebra, Geometry and Analysis
of Bicomplex Numbers
M. Elena Luna-Elizarrarás Michael Shapiro
Escuela Sup. de Física y Matemáticas Escuela Sup. de Física y Matemáticas
Instituto Politécnico Nacional Instituto Politécnico Nacional
Mexico City, Mexico Mexico City, Mexico
Introduction 1
v
vi Contents
Bibliography 219
Index 226
Introduction
The best known extension of the field of complex numbers to the four-dimensional
setting is the skew field of quaternions, introduced by W.R. Hamilton in 1844,
[36], [37]. Quaternions arise by considering three imaginary units, i, j, k that an-
ticommute and such that ij = k. The beauty of the theory of quaternions is that
they form a field, where all the customary operations can be accomplished. Their
blemish, if one can use this word, is the loss of commutativity. While from a purely
algebraic point of view, the lack of commutativity is not such a terrible problem,
it does create many difficulties when one tries to extend to quaternions the fecund
theory of holomorphic functions of one complex variable. Within this context, one
should at least point out that several successful theories exist for holomorphicity
in the quaternionic setting. Among those the notion of Fueter regularity (see for
example Fueter’s own work [27], or [97] for a modern treatment), and the theory
of slice regular functions, originally introduced in [30], and fully developed in [31].
References [97] and [31] contain various quaternionic analogues of the bicomplex
results presented in this book.
It is for this reason that it is not unreasonable to consider whether a four-
dimensional algebra, containing C as a subalgebra, can be introduced in a way that
preserves commutativity. Not surprisingly, this can be done by simply considering
two imaginary units i, j, introducing k = ij (as in the quaternionic case) but now
imposing that ij = ji. This turns k into what is known as a hyperbolic imaginary
unit, i.e., an element such that k2 = 1. As far as we know, the first time that these
objects were introduced was almost contemporary with Hamilton’s construction,
and in fact J.Cockle wrote, in 1848, a series of papers in which he introduced a
new algebra that he called the algebra of tessarines, [15, 16, 17, 18]. Cockle’s work
was certainly stimulated by Hamilton’s and he was the first to use tessarines to
isolate the hyperbolic trigonometric series as components of the exponential series
(we will show how this is done later on in Chapter 6). Not surprisingly, Cockle
immediately realized that there was a price to be paid for commutativity in four
dimensions, and the price was the existence of zero-divisors. This discovery led
him to call such numbers impossibles, and the theory had no further significant
development for a while.
It was only in 1892 that the mathematician Corrado Segre, inspired by the
work of Hamilton and Clifford, introduced what he called bicomplex numbers in
[82], their algebra being equivalent to the algebra of tessarines. It was in his original
1 + ij 1 − ij
papers that Segre noticed that the elements and are idempotents
2 2
and play a central role in the theory of bicomplex numbers. Following Segre, a few
other mathematicians, in particular Spampinato [88, 89] and Scorza Dragoni [83],
developed the first rudiments of a function theory on bicomplex numbers.
The next major push in the study of bicomplex analysis was the work of
J.D.Riley, who in 1953 published his doctoral dissertation [57] in which he further
developed the theory of functions of bicomplex variables. But the most important
contribution was undoubtedly the work of G. B. Price, [56], where the theory of
holomorphic functions of a bicomplex variable (as well as multicomplex variables)
is widely developed. Until this monograph, the work of G. B. Price had to be
regarded as the foundational work in this theory.
In recent years, however, there has been a resurgence of interest in the study
of holomorphic functions on one and several bicomplex variables, as well as a sig-
nificant interest in developing functional analysis on spaces that have a structure
of modules over the ring of bicomplex numbers. Without any pretense of com-
pleteness, we refer in this book to [2, 12, 13, 14, 19, 20, 29, 32, 34, 45, 59, 61, 62,
63, 65, 96]. Most of this new work indicates a need for the development of the
foundations of the theory of holomorphy on the ring of bicomplex numbers, that
better expresses the similarities, and differences, with the classical theory of one
complex variable.
This is the explicit and intentional purpose of this book, which we have writ-
ten as an elementary, yet comprehensive, introduction to the algebra, geometry,
and analysis of bicomplex numbers.
We describe now the structure of this work. Chapter 1 introduces the funda-
mental properties of bicomplex numbers, their definitions, and the different ways
in which they can be written. In particular, we show how hyperbolic numbers can
be recognized inside the set of bicomplex numbers. The algebraic structure of this
set is described in detail in the next chapter, where we define linear spaces and
modules on BC and we introduce a partial order on the set of hyperbolic num-
bers. Maybe the most important contribution in this chapter is the definition of
a hyperbolic-valued norm on the ring of bicomplex numbers. This norm will have
great importance in all future applications of bicomplex numbers. In Chapter 3
we move into geometry, and we spend considerable time in discussing how to visu-
alize the 4-dimensional geometry of bicomplex numbers. We also discuss the way
in which the trigonometric representation of complex numbers can be extended to
the ring of bicomplex numbers. In Chapter 4 we remain in the geometric realm
and discuss lines in BC; in particular we study real, complex, and hyperbolic lines
in BC. We then extend this analysis to the study of hyperbolic and complex curves
in BC. With Chapter 5 we abandon geometry and begin the study of analysis of
bicomplex functions. We discuss here the notion of limit in the bicomplex con-
text, which will be necessary when we study holomorphy in the bicomplex setting.
Introduction 3
Acknowledgments. This work has been made possible by frequent exchanges be-
tween the Instituto Politécnico Nacional in Mexico, D.F., and Chapman University
in Orange, California. The authors express their gratitude to these institutions for
facilitating their collaboration. A very special thank you goes to M. J. C. Robles–
Casimiro, who skillfully prepared all the drawings that are included in this volume.
Chapter 1
Z · W = W · Z, Z · (W · Y ) = (Z · W ) · Y,
Z · (W + Y ) = Z · W + Z · Y ,
inherits all the algebraic definitions, operations and properties from BC.
1.2. Versatility of different writings of bicomplex numbers 7
will be especially useful later. We will call its elements “non-negative hyperbolic
numbers”; the set
D+ \ {0} = {x + ky x2 − y 2 ≥ 0, x > 0 }
will be called the set of “positive hyperbolic numbers”. Such definitions of “non-
negativeness” and of “positiveness” for hyperbolic numbers do not look intuitively
clear but later on we will give them other descriptions clarifying the reason for such
names. It turns out that the non-negative hyperbolic numbers play with respect to
all hyperbolic numbers a role deeply similar to that of real non-negative numbers
with respect to all real numbers.
The set
D− := {x + ky x2 − y 2 ≥ 0, x ≤ 0 } = {z − z ∈ D+ }
will bear the name of non-positive hyperbolic numbers; and of course the set
D− \ {0} = {x + ky x2 − y 2 ≥ 0, x < 0 }
is the set of negative hyperbolic numbers. Clearly, there are hyperbolic numbers
which are neither non-negative nor non-positive.
(i’) Z = ζ1 − i ζ2 = z1 − i z2 = w1 + j w2 = w1 − k w2 = ω1 − k ω2
= x 1 − i y1 + j x 2 − k y2 ;
1.4. Moduli of bicomplex numbers 9
† ∗
(Z + W ) = Z + W , (Z + W ) = Z † + W † , (Z + W ) = Z ∗ + W ∗ (1.10)
† † ∗
Z = Z, Z = Z, (Z ∗ ) = Z (1.11)
† † † ∗ ∗ ∗
(Z · W ) = Z · W , (Z · W ) = Z · W , (Z · W ) = Z · W . (1.12)
= w12 − w22
= |ω1 |2 − |ω2 |2 − 2 Im ω1† ω2 i ∈ C(i);
• |Z|2j := Z · Z = |z1 |2 − |z2 |2 + 2 Re (z1 z 2 )j
= ζ12 + ζ22
= |z1 |2hyp − |z2 |2hyp + z1 z2 + z1 z2 j
= |w1 |2hyp + |w2 |2hyp + w1 w2 − w1 w2 i
= |w1 |2 − |w2 |2 + (w2 w1 − w1 w2 ) k
= ω12 − ω22 ∈ C(j);
10 Chapter 1. The Bicomplex Numbers
• |Z|2k := Z · Z ∗ = |z1 |2 + |z2 |2 − 2 Im (z1 z 2 )k
= |ζ1 |2 + |ζ2 |2 − 2 Im (ζ1 ζ2† )k
= z21 + z22
= w21 + w22
= |w1 |2 + |w2 |2 + (w2 w1 + w1 w2 ) k
= |ω1 |2 + |ω2 |2 + (ω1 ω2∗ + ω2 ω1∗ ) k ∈ D,
where for a complex number z (in C(i) or C(j)) we denote by |z| its usual modulus
and for a hyperbolic number z = a + kb we use the notation |z|2hyp = a2 − b2 .
Unlike what happens in the complex case, these moduli are not R+ -valued.
The first two moduli are complex-valued (in C(i) and C( j) respectively), while the
last one is hyperbolic-valued.
√
The value of |Z|i = Z · Z † , being the square root of a complex number, is
†
√ for the complex number z = Z · Z , if z
determined by the following convention:
non-negative real number, then z denotes its non-negative value; otherwise,
is a √
the z denotes the value of the square root of z in the upper half-plane. In many
standard references, this latter one is also called the “principal” square root of z.
Although in general |Z|i is a C(i)-complex number, nevertheless if Z is in
C( j), then its C(i)-complex modulus |Z|i coincides with the usual modulus of the
complex number ζ1 = x1 + jx2 : since z1 = x1 + i0, z2 = x2 + i0, then
|Z|i = x21 + x22 = |ζ1 | .
Hence the restriction of the quadratic form z12 + z22 onto the real two-dimensional
plane C(j) determines the usual Euclidean structure on this plane.
We make similar conventions for the C(j)-valued modulus
|Z|j = Z · Z.
ZZ = ζ12 + ζ22
onto the real two-dimensional plane C(i) determines the usual Euclidean structure
on this plane.
We observe here a kind of a “dual” relation between the two types of complex
moduli and the respective complex numbers: if Z = z1 ∈ C(i), then
|Z|i = |z1 |i = z12 ,
1.4. Moduli of bicomplex numbers 11
which, in general, is not equal to |z1 | but is equal to z1 or −z1 ; but somewhat
paradoxically, if Z = ζ1 ∈ C(j), then |Z|i = |ζ1 |.
Similarly, the |Z|j of C(j)-numbers, Z = ζ1 , is |Z|j = |ζ1 |j = ζ12 , meanwhile
if Z = z1 ∈ C(i), then |z1 |j = |z1 |. We will refer to |Z|i , |Z|j as the C(i)- and
C(j)-valued moduli of the bicomplex number Z respectively.
The last modulus introduced has its square, | · |2k , which is hyperbolic-valued,
and later we will show that the modulus itself can be chosen hyperbolic-valued.
For its square the following holds:
|Z|2k = |z1 |2 + |z2 |2 + k (−2 Im(z1 z 2 ) ) =: x + k y ,
where x and y satisfy x2 − y 2 ≥ 0 (this is proved using the fact that | Im(z1 z 2 ) | ≤
|z1 | · |z2 |). Thus |Z|2k ∈ D+ .
We will specify the value of the square root of a hyperbolic number later on.
Although these moduli are not real-valued, nevertheless they preserve, for-
tunately, an important property related with the multiplication; specifically, we
have:
or as
C2 (j) = { (ζ1 , ζ2 ) | ζ1 + i ζ2 ∈ BC } ,
or as
R4 = { (x1 , y1 , x2 , y2 ) | (x1 + i y1 ) + j(x2 + i y2 ) ∈ BC } .
The Euclidean norm |Z| is related with the properties of bicomplex numbers via
the D+ -valued modulus:
|Z| = |z1 |2 + |z2 |2 = |ζ1 |2 + |ζ2 |2 = Re (|Z|2k ) = x21 + y12 + x22 + y22 ,
Z Z∗
Z −1 = = .
|Z|2j |Z|2k
Z† z1 − jz2
Z −1 = = 2 .
z12 + z22 z1 + z22
If both z1 and z2 are non-zero but the sum z12 + z22 = 0, then the corresponding
bicomplex number Z = z1 + jz2 is a zero-divisor. This is equivalent to z12 = −z22 ,
i.e.,
z1 = ±iz2 , (1.17)
and thus all zero-divisors in BC are of the form:
At first sight, we have something quite different from (1.17). Note however that
Re(ζ1 ζ2† ) is the Euclidean inner product in R2 , hence (1.19) means that ζ1 and ζ2
are orthogonal in C(j) and with the same magnitude (i.e., with the same modulus
of complex numbers), and thus
ζ1 = ± j ζ2 .
Z = ζ1 ± ijζ1 = ζ1 (1 ± i j) (1.20)
Z · Z = ζ12 + ζ22 = 0
14 Chapter 1. The Bicomplex Numbers
which uses yet another conjugation, not the †-conjugation but the bar-conjugation.
It is possible to give several other descriptions of the set of zero-divisors
using all the three conjugations as well as formulas (1.3)–(1.9). This we leave as
an exercise to the reader.
We denote the set of all zero-divisors in BC by S, and we set S0 := S ∪ {0}.
We can summarize this discussion as follows.
Theorem 1.5.1. Let Z = 0, then the following are equivalent.
1. The bicomplex number Z is invertible.
2. Z is not a zero-divisor.
3. Z · Z † = 0.
4. Z · Z = 0.
5. Z · Z ∗ ∈ S0 .
6. |Z|i = 0.
7. |Z|j = 0.
8. |Z|k ∈ S0 .
9. If Z is given as Z = z1 + jz2 , then z12 + z22 = 0.
10. If Z is given as Z = ζ1 + iζ2 , then ζ12 + ζ22 = 0.
Since BC is a ring (we will comment on this with more detail in the next
chapter) it is worth to single out the equivalence between (1) and (2) in Theorem
1.5.1. Indeed, in a general ring, the set of non-zero elements which are not zero-
divisors is a different set from the set of invertible elements; from this point of
view BC is a remarkable exception.
Of course the above Theorem allows us to give immediately a “dual” char-
acterization of the set of zero-divisors.
Corollary 1.5.2. Let Z = 0, then the following are equivalent.
1. Z is not invertible.
2. Z is a zero-divisor.
3. Z · Z † = 0 = Z · Z.
4. Z · Z ∗ ∈ S0 .
5. |Z|i = 0 = |Z|j .
6. |Z|k ∈ S0 .
7. If Z is given as Z = z1 + jz2 , then z12 + z22 = 0.
8. If Z is given as Z = ζ1 + iζ2 , then ζ12 + ζ22 = 0.
1.6. Idempotent representations of bicomplex numbers 15
e2 = e, (e† )2 = e†
e + e† = 1, e − e† = ij .
i e = −j e, i e† = j e† ,
k e = e, ke† = −e† . (1.21)
The next property has no analogs for complex numbers, and it exemplifies
one of the deepest peculiarities of the set of bicomplex numbers. For any bicomplex
number Z = z1 + jz2 ∈ BC we have:
z1 − iz2 + z1 + iz2 z2 + iz1 + z2 − iz1
Z = z1 + jz2 = +j
2 2
z1 − iz2 z1 + iz2 z1 − iz2 z1 + iz2
= + + ij − ij
2 2 2 2
1 + ij 1 − ij
= (z1 − iz2 ) + (z1 + iz2 ) ,
2 2
that is,
Z = β 1 e + β 2 e† , (1.22)
where β1 := z1 −iz2 and β2 := z1 +iz2 are complex numbers in C(i). Formula (1.22)
is called the C(i)-idempotent representation of the bicomplex number Z.
It is obvious that since β1 and β2 are both in C(i), then β1 e + β2 e† = 0 if and
only if β1 = 0 = β2 . This implies that the above idempotent representation of the
bicomplex number Z is unique: indeed, assume that Z = 0 has two idempotent
representations, say,
Z = β1 e + β2 e† = β1 e + β2 e† ,
16 Chapter 1. The Bicomplex Numbers
Z + W = (β1 + ν1 ) e + (β2 + ν2 ) e† ,
Z · W = (β1 ν1 ) e + (β2 ν2 ) e† ,
Z n = β1n e + β2n e† .
The proof of the formulas in the proposition above relies simply on the rather
specific properties of the numbers e and e† . For example, let us prove the second
property:
Z · W = β 1 e + β 2 e† · ν 1 e + ν 2 e †
= β1 e · ν1 e + β1 e · ν2 e† + β2 e† · ν1 e + β2 e† · ν2 e†
= β 1 ν 1 · e + β 1 ν 2 · 0 + β 2 ν 1 · 0 + β 2 ν 2 · e†
= β 1 ν 1 · e + β 2 ν 2 · e† .
We used the fact that e and e† are idempotents, i.e., each of them squares to itself,
and that their product is zero.
We showed after formula (1.22) that the coefficients β1 and β2 of the idem-
potent representation are uniquely defined complex numbers. But this refers to
the complex numbers in C(i), and the paradoxical nature of the idempotents e
and e† manifests itself as follows.
Take a bicomplex number Z written in the form Z = ζ1 + i ζ2 , with ζ1 , ζ2 ∈
C(j). Then a direct computation shows:
Z = β 1 e + β 2 e † = α 1 e + α 2 e† . (1.24)
Let us find out which is the relation between them. One has that
eZ = β1 e = α1 e
and
e† Z = β 2 e† = α 2 e † ,
1.6. Idempotent representations of bicomplex numbers 17
thus the authentic uniqueness consists of the fact that not the coefficients β1 and
α1 (or β2 and α2 ) are equal, but the products β1 e and α1 e (or β2 e† and α2 e† ) are
equal respectively. What is more, β1 e = α1 e is equivalent to (β1 − α1 )e = 0, but
since e is a zero-divisor, then β1 − α1 is also a zero-divisor, that is, β1 − α1 = A· e† ,
where A can be chosen either in C(i) or in C(j). The latter is justified with the
following reasoning. Take β1 , β2 to be β1 = c1 + id1 , β2 = c2 + id2 , then
Z = β1 e + β2 e† = (c1 + i d1 ) e + (c2 + i d2 ) e†
= c 1 e − j d 1 e + c 2 e† + j d 2 e†
= e · (c1 − j d1 ) + e† · (c2 + j d2 )
= e · α 1 + e† · α 2 ,
where α1 = c1 − j d1 , α2 = c2 + j d2 ; thus
β1 − α1 = c1 + i d1 − c1 + j d1 = d1 (i + j)
= i d1 (1 − i j)
= 2 d1 i e† = 2 d1 j e† .
Z = (1 + i) + j (3 − 2i) =: z1 + j z2 .
Z = (1 + 3j) + i (1 − 2j) =: ζ1 + i ζ2 .
β1 − α1 = d1 (i + j).
Since
β1 − α1 = −2 (i + j) = −4 i e† = −4 j e† ,
one obtains d1 = −2, which coincides with the value of d1 in this example.
18 Chapter 1. The Bicomplex Numbers
Let us see now how the conjugations and moduli manifest themselves in
idempotent representations. Take Z = β1 e + β2 e† = α1 e + α2 e† , with β1 and β2
in C(i), α1 and α2 in C( j). Then it is immediate to see that
Z = β 2 e + β 1 e† = α 2 e + α 1 e † ;
Z † = β2 e + β1 e† = α2† e + α1† e† ;
Z ∗ = β 1 e + β 2 e† = α1† e + α2† e† .
Hence, the squares of all the three moduli become:
2
Z = Z · Z
j
= β 1 e + β 2 e† · β 2 e + β 1 e†
= β 1 β 2 e + β 1 β 2 e†
= α 1 e + α 2 e† · α 2 e + α 1 e †
= α1 α2 e + α1 α2 e† = α1 α2 ∈ C( j);
2
Z = Z · Z †
i
= β 1 e + β 2 e† · β 2 e + β 1 e†
= β 1 β 2 e + β 1 β 2 e† = β 1 β 2
= α1 e + α2 e† · α2† e + α1† e†
†
= α1 α2† e + α1 α2† e† ∈ C(i);
2
Z = Z · Z ∗
k
= β 1 e + β 2 e† · β 1 e + β 2 e †
2 2
= β 1 β 1 e + β 2 β 2 e† = β 1 e + β 2 e†
= α1 e + α2 e† · α1† e + α2† e†
2 2
= α1 α1† e + α2 α2† e† = α1 e + α2 e† ∈ D+ .
Observe that in the formulas for |Z|2k the idempotent coefficients are non-
negative real numbers and we will see soon that this is a characteristic property
of non-negative hyperbolic numbers. Observe also that given Z = β1 e + β2 e† =
α1 e + α2 e† with β1 , β2 in C(i) and α1 , α2 in C( j), then
1 1
|Z| = √ |β1 |2 + |β2 |2 = √ |α1 |2 + |α2 |2 .
2 2
We can characterize now the invertibility of bicomplex numbers in terms of
the idempotent representations.
1.6. Idempotent representations of bicomplex numbers 19
One can ask if there are more idempotents in BC, not only e and e† (of course
the trivial idempotents 0 and 1 do not count). Assume that a bicomplex number
Z = β1 e + β2 e† , with β1 and β2 being complex numbers either in C(i) or C( j), is
an idempotent: Z 2 = Z. Then
β12 e + β22 e† = β1 e + β2 e†
and
β12 = β1 and β22 = β2 ,
which gives:
β1 ∈ {0, 1}, β2 ∈ {0, 1}.
Hence, combining all possible choices we have at most four candidates for idem-
potents in BC:
Z1 = 0 · e + 0 · e† = 0,
20 Chapter 1. The Bicomplex Numbers
Z2 = 1 · e + 1 · e† = 1,
Z3 = 1 · e + 0 · e† = e,
Z 4 = 0 · e + 1 · e† = e† .
Thus, one concludes that e and e† are the only non-trivial idempotents in BC.
Remark 1.6.7. The formulas
Z = β 1 e + β 2 e† and Z † = β 2 e + β 1 e† ,
β1 = β1 e + β1 e† = Ze + Z † e† ;
β2 = β2 e† + β2 e = Ze† + Z † e.
γ1 = γ1 e + γ1 e† = Ze + Ze† ;
γ2 = γ2 e† + γ2 e = Ze + Ze† .
z · z = x2 − y 2 ∈ R , (1.25)
|z|2hyp := x2 − y 2 ,
non-zero, but the product is zero: z·z = 0. All zero-divisors in D are characterized
by x2 = y 2 , i.e., x = ±y, thus they are of the form
z = λ(1 ± k)
z = se + te . (1.27)
Observe that
|z|2hyp = x2 − y 2 = (x + y)(x − y) = st .
Let us show now how these properties are related with their bicomplex an-
tecedents. We are interested in bicomplex numbers Z = z1 + j z2 with Im(z1 ) =
0 = Re(z2 ), that is, our hyperbolic numbers are of the form z = x1 + ij y2 and the
hyperbolic unit is k = ij. Then the -conjugation operation is consistent with the
bicomplex conjugations † and bar in the following way:
†
z = ((x1 + i0) + j(0 + iy2 )) = ((x1 + i0) + j(0 + iy2 )) = x1 − ky2 .
For this reason, from this point on we will not write the hyperbolic conjugate of
e as e anymore, but we will use the bicomplex notation e† .
For a general bicomplex number, the three moduli have been defined in Sec-
tion 1.4. Let us see what happens if they are evaluated on a generic hyperbolic
number z = x1 + ky2 . Considering it as z = z1 + jz2 := (x1 + i0) + j(0 + iy2 ) ∈ BC,
we have:
This is not the case of the third modulus: the hyperbolic-valued modulus of Z = z
is different than the intrinsic modulus of z. Indeed, we have:
|z|2k = Z · Z ∗ = Z · Z = Z 2 = z2 . (1.29)
In (1.28) we have a relation between the squares of the three moduli |z|i , |z|j
and |z|hyp for hyperbolic numbers. The question now is how to define the modulus
|z|hyp itself, which obviously should be defined as the square root of x21 − y22 . Note
that some authors consider the non-negative values of x21 − y22 only.
It is instructive to analyze the situation more rigorously and to understand
if we have other options for choosing an appropriate value of the intrinsic modu-
lus. Although we work here with hyperbolic numbers, at the same time one can
think about bicomplex numbers also as of possible values of the square roots of a
hyperbolic number. So let us consider the solutions in BC of the equation Z 2 = R
for a given real number R. Write Z = β1 e + β2 e† , then the equation Z 2 = R is
equivalent to
β12 e + β22 e† = Re + Re†
which is equivalent to
β12 = R; β22 = R.
If R = 0, then the only solution is Z = 0. If R is positive, then
√ √
β1 = ± R; β2 = ± R,
These are all the solutions in BC, and they are real or hyperbolic numbers.
If R is negative, then one gets:
√ √
β1 = ±i −R, β2 = ±i −R
Thus, for R < 0 the equation Z 2 = R has four solutions none of which is a
hyperbolic number; two of them are complex numbers in C(i) and the remaining
two are complex numbers in C( j).
Returning to the intrinsic modulus |z|hyp of a hyperbolic number z we see
that in case x21 − y22 > 0 this modulus can be taken as a positive real number
1.7. Hyperbolic numbers inside bicomplex numbers 23
x21 − y22 or even as a hyperbolic number ±k x21 − y22 . But if x21 − y22 < 0, then
there are no solutions in D, the candidates should be taken as complex (in C(i) or
in C( j)) numbers.
2 2
It is instructive to note that in case x1 − y2 > 0 the positive real number
2 2
x1 − y2 coincides with the equal values of |z|i and |z|j as defined in Section 1.4.
When x21 − y22 < 0, then |z|hyp can be chosen either as |z|i ∈ C(i) or as
|z|j ∈ C( j) (recall that we have agreed to take, in both cases, the value of the
square root which is in the upper half plane); as formula (1.29) shows, it cannot
be chosen as |z|k .
Recall also that we have defined the set D+ of non-negative hyperbolic numbers
as
D+ = x + ky x2 − y 2 ≥ 0, x ≥ 0 .
The first of the defining inequalities gives the two systems:
x − y ≥ 0, x − y ≤ 0,
or
x + y ≥ 0, x + y ≤ 0,
but the condition x ≥ 0 eliminates the second system; hence, the set D+ can be
described as
D+ = x + ky x ≥ 0; |y| ≤ x ,
or as
D+ = {νe + μe† ν, μ ≥ 0 }.
Thus positive hyperbolic numbers are those hyperbolic numbers whose both idem-
potent components are non-negative, that somehow explains the origin of the
name.
In Fig. 1.7.1 the points (x, y) correspond to the hyperbolic numbers z =
x + ky. One sees that, geometrically, the hyperbolic positive numbers are situated
in the quarter plane denoted by D+ . The quarter plane symmetric to it with respect
to the origin corresponds to the negative hyperbolic numbers. The other points
correspond to those hyperbolic numbers which cannot be called either positive or
negative.
24 Chapter 1. The Bicomplex Numbers
2k
1
D− D+
e
O 1 x
2
− 12 k e†
D− = x + ky x ≤ 0; |y| ≤ |x| ,
or equivalently
D− = {νe + μe† ν, μ ≤ 0 }.
We will say sometimes that the hyperbolic number z = νe + μe† is semi-positive
if one of the coefficients μ and ν is positive and the other is zero.
We mentioned already that D+ plays an analogous role as non-negative real
numbers, and now we illustrate this by computing the square roots of a hyperbolic
number in D+ . Take z ∈ D+ , then z = μe + νe† with μ, ν ∈ R+ ∪ {0}, and it is
easy to see that all the four hyperbolic numbers
√ √
± μ e ± ν e†
√
√ † to z, but only one of them is a non-negative hyperbolic number: μ e +
square
νe .
We are now in a position to define the meaning of the symbol |Z|k for any
bicomplex number Z = β1 e + β2 e† . Indeed, we have obtained that |Z|2k = |β1 |2 e +
|β2 |2 e† which is a non-negative hyperbolic number, hence the modulus |Z|k can
1.8. The Euclidean norm and the product of bicomplex numbers 25
be taken as
|Z|k := |β1 |e + |β2 |e† ∈ D+ .
We will come back to this in the next chapter considering the notion of BC as a
bicomplex normed module where the norm will be D+ -valued. Meanwhile we can
complement the above reasoning solving the equation
|z|k = w
z = ±γ2 e† or z = ±γ1 e ,
respectively.
• If w is positive but not semi-positive: γ1 > 0 and γ2 > 0, then all four
solutions are
z = ±γ1 e ± γ2 e† .
|Z · U | =
|Z| · |U | .
More precisely,
|Z · U |2 = |Z|2 · |U |2 + 4 x1 y2 Re(i u1 u2 ).
where we used the fact that the Euclidean norm of a complex number (both in
C(i) and in C( j)), seen as a bicomplex number, coincides with its modulus.
Take now Z = x1 + jx2 = (x1 − i x2 ) e + (x1 + i x2 ) e† ∈ C(j), then
|Z · U |2 = | (x1 − i x2 ) e + (x1 + i x2 ) e†
· (u1 − i u2 ) e + (u1 + i u2 ) e† |
= |(x1 − i x2 )(u1 − i u2 )e + (x1 + i x2 )(u1 + i u2 )e† |2
1
= |x1 − ix2 |2 · |u1 − iu2 |2 + |x1 + ix2 |2 · |u1 + iu2 |2
2
= |Z|2 · |U |2 .
|z1 − i z2 |2 = |z1 + i z2 |2
which is equivalent to
z1 · z 2 = λ ∈ R.
The following cases arise:
(1) if λ = 0, then Z is in C(i), or Z = j z2 ∈ j · C(i), or both; the last means that
Z = 0;
(2) if λ = 0, then
λ
z1 = · z2 ,
|z2 |2
λ
i.e., Z = z2 + j and Z becomes the product of a C(i)-complex num-
|z2 |2
ber and a C( j)-complex number.
Let us show that the reciprocal is also true. Take a = a1 + i a2 , b = b1 + j b2 ,
where a1 , a2 , b1 , b2 are real numbers, and set Z := a · b = a · b1 + j a · b2 , then
Z = (a b1 − i a b2 ) · e + (a b1 + i a b2 ) · e†
= a · (b1 − i b2 ) · e + a (b1 + i b2 ) · e† =: β1 e + β2 e†
and thus
1 √
≤ 2 · |Z −1 |.
|Z|
We ask now for which class of bicomplex numbers the “conventional” formula
1
|Z −1 | = (1.31)
|Z|
holds? The answer follows from the conditions ensuring the equality |Z · W | =
|Z| · |W |, in which we can take W = Z −1 , thus obtaining that (1.31) holds if and
only if Z is a product of a complex number in C(i) by a complex number in C( j)
or, equivalently, if and only if the Euclidean norm of Z coincides with the modulus
of any of its idempotent components.
Chapter 2
Z (Z1 + Z2 ) = Z Z1 + Z Z2 . (2.1)
In this case the bicomplex numbers 1 and j are mapped into the canonical basis of
C2 (i). Composing this isomorphism and the inverse of the previous one, we have
the following isomorphism between R4 and C2 (i):
Seeing now BC as a C( j)-linear space and using (1.4), we have the following
isomorphism:
BC
Z = ζ1 + iζ2 −→ (ζ1 , ζ2 ) ∈ C2 ( j). (2.5)
This isomorphism sends the bicomplex numbers 1 and i into the canonical basis
in C2 ( j) and it induces the following isomorphism (of real linear spaces) between
R4 and C2 ( j):
Obviously the isomorphisms (2.4) and (2.6) are different: this shows once
again that inside BC the “complex sets” C2 (i) and C2 ( j) play distinct roles.
2.2. Linear spaces and modules in BC 31
This isomorphism arose implicitly in the first chapter when we compared the
two idempotent representations, one with coefficients in C(i) and another with
coefficients in C( j):
Z = β1 e + β2 e† = (c1 + i d1 ) e + (c2 + i d2 ) e†
(2.11)
= α1 e + α2 e† = (c1 − j d1 ) e + (c2 + j d2 ) e† ,
†
which shows that α1 = (ϕ(β1 )) and α2 = ϕ(β2 ).
Following the line of defining isomorphisms between BC and the complex
spaces C2 (i) and C2 ( j), we will see that the idempotent representations suggest
two more complex linear spaces isomorphisms. We need first the following propo-
sition.
Proposition 2.2.1. The zero-divisors
1 + ij 1 − ij
e= and e† =
2 2
are linearly independent in BC when it is seen as a C(i)-linear space or as a
C( j)-linear space.
Proof. Using the isomorphism (2.3), the bicomplex numbers e and e† are mapped
as follows:
1 i 1
e −→ , = (1, i) ,
2 2 2
1 i 1
e† −→ ,− = (1, −i) ,
2 2 2
and considering the equation
Again, the relations between (2.3) and (2.12) as well as between (2.5) and
(2.13) are given through the change of basis from the canonical ones to the basis
1 i 1 −i
, , ,
2 2 2 2
2.3. Algebra structures in BC 33
and
It follows directly from the properties of the bicomplex multiplication that (2.17)
endows R4 with the structure of a commutative real algebra. Moreover, the iso-
morphism (2.2) extends up to a real algebras isomorphism.
34 Chapter 2. Algebraic Structures of the Set of Bicomplex Numbers
It is well known that there are not too many “reasonably good” multipli-
cations in R4 . The bicomplex multiplication given by (2.17) is one of these few
options and one can compare it with the multiplication generated by quaternions.
The next step is to consider both complex algebras. The isomorphisms (2.3),
(2.7) and (2.12) give us three candidates for introducing a multiplication on the
C(i)-linear space C2 (i):
w 1 = z1 , θ 1 = p1 ,
and
w2 = −i z2 , θ2 = −i p2 ,
thus,
The case of the complex linear space C2 ( j) is treated in exactly the same
way.
Now taking into account that BC is a module over D and over itself, and that
it is also a ring, we conclude that the set BC is a D-algebra and a BC-algebra. As
such, BC is isomorphic to the D-algebra D2 and the isomorphisms are given by
formulas (2.15) and (2.16).
2.4. Matrix representations of bicomplex numbers 35
In particular, the matrices of this form commute under multiplication and any of
them, but zero, has its inverse. Under this isomorphism the imaginary unit i is
represented by the matrix
0 −1
I := .
1 0
Take z = x + iy, then
x −y 1 0 0 −1
φC (z) = =x +y = xI2 + yI .
y x 0 1 1 0
The square of the modulus of a complex number z coincides with det φC (z).
In terms of representation theory the mapping φC is called a representation
of the field C into a subset of the set of 2 × 2 real matrices.
A similar reasoning applies to the ring BC. The mapping
z1 −z2
φC(i) : Z = z1 + jz2 ∈ BC −→ (2.18)
z2 z1
turns out to be an isomorphism (of rings) between BC and the set of matrices
z1 −z2
z1 , z2 ∈ C(i) .
z2 z1
y2 x 2 y1 x1
Every 4 × 4 matrix determines a linear (more exactly, a real linear) trans-
formation on R4 . Of course, not all of them remain BC-linear when R4 is seen as
BC. Those matrices which represent BC-linear mappings are of the form φR (Z).
Notice also that at this stage, we are considering again the following identi-
fication between BC and R4 :
Z = x1 + iy1 + jx2 + ky2 ←→ (x1 , y1 , x2 , y2 ) .
This “forces” the following two identifications between BC and C2 :
Z = z1 + jz2 ←→ (z1 , z2 ) = (x1 + iy1 , x2 + iy2 ) ∈ C2 (i) ←→ (x1 , y1 , x2 , y2 )
and
Z = ζ1 + iζ2 ←→ (ζ1 , ζ2 ) = (x1 + jx2 , y1 + jy2 ) ∈ C2 (j) ←→ (x1 , y1 , x2 , y2 ) .
Consider now an R-linear mapping T : R4 → R4 which represents also a C(i)-linear
mapping, then the matrix of T is of the form
⎛ ⎞
a −b c −d
⎜ b a d c ⎟
⎜ ⎟
⎝ −m u −v ⎠ ,
m v u
meanwhile, if T represents a C(j)-linear mapping, then its matrix is of the form
⎛ ⎞
A B −E −F
⎜ C D −G −H ⎟
⎜ ⎟.
⎝ E F A B ⎠
G H C D
2.5. Bilinear forms and inner products 37
It is clear that, as one should expect, the matrix φR (Z) represents both a C(i)-
linear mapping and a C(j)-linear one.
BR (x, y) := x · y .
Q := BR (x, x) = x2
BC,R (z, w) := xu + yv ,
and
QC,2 (z) := BC,2 (z, z) := |z|2 = x2 + y 2 ,
where again QC,2 (z) = QC,R (z) is the square of the Euclidean metric on C. The
forms BC,1 and QC,1 are employed widely in different areas of mathematics, but
BC,1 is not called usually an inner product and QC,1 does not define any metric in
the classical sense.
38 Chapter 2. Algebraic Structures of the Set of Bicomplex Numbers
Let us extend these ideas onto the bicomplex context. Starting with the real
structure on BC, we see it as R4 = {(x1 , y1 , x2 , y2 ) = Z} and thus we endow it
with the (real) bilinear form
BBC,R (Z, W ) := x1 u1 + y1 v1 + x2 u2 + y2 v2 ,
where we notice that the last one coincides with QBC;i,2 (Z), and both are equal
to the square of the Euclidean metric.
When BC is interpreted as D2 , the situation is different. Of course, one sets:
1
BBC;D,1 (Z, W ) := z1 w1 + z2 w2 = (ZW ∗ + Z ∗ W )
2
and
1 1
BBC;D,2 (Z, W ) := z1 w1 + z2 w2 = (ZW † + Z ∗ W ) = (ZW + Z ∗ W † ) ,
2 2
imitating the previous situations, but now both forms take values in D, not in C
or R. Note also that the first of them is hyperbolic bilinear, and the second one
can be called hyperbolic sesquilinear; what is more, setting
BBC (Z, W ) := Z · W
The corresponding quadratic forms coincide with the three “moduli” previously
introduced, which take complex or hyperbolic values, making the geometric aspect
even more complicated than the above described case of the D-module BC = D2 .
Of course, this makes both cases even more interesting and intriguing.
Let us consider again the R-valued quadratic form QC,2 (z) = x2 + y 2 ; since
it coincides with BC,2 (z, z), then QC,2 enjoys the factorization
This identity can be seen as one of the reasons for the necessity of introducing
complex numbers: if one wants to factorize QC,2 (z) (which is a real-valued and
positive definite quadratic form; thus, in particular, the set of its values is R-
one-dimensional) into the product of two linear forms which should be real two-
dimensional, then the imaginary unit i emerges forcedly and generates the whole
set C.
A very similar idea is related with the bicomplex numbers. Consider the
C(i)-valued quadratic form QBC,i,1 (Z) = z12 + z22 . We know that it factorizes into
where the set of the values of QBC,i,1 is C(i)-one-dimensional but the factors are
already C(i)-two-dimensional. Thus, the C(i)-algebra BC arises from a complex
quadratic form in the same way as the real algebra C arises from a real quadratic
form. Notice that the requirement for the factors to be C(i)-two-dimensional, not
one-dimensional, is crucial since without it one has an obvious factorization
for all z1 and z2 in C(i). Hence, for all z1 and z2 it holds that
which is equivalent to
ac = 1; bd = 1; ad + bc = 0.
c−1 d + d−1 c = 0,
i.e.,
2
c−1 d = −1.
Therefore, denoting j := c−1 d we have that j2 = −1 and j−1 = −j; the factoriza-
tion becomes
z12 + z22 = (z1 + jz2 )(z1 − jz2 ).
Thus, the complex algebra we are looking for should be generated by 1 and by a
new element j = ±i, and we have arrived exactly at BC.
2.6. A partial order on the set of hyperbolic numbers 41
In the same way we can begin with the C( j)-valued quadratic form ζ12 + ζ22
and get the same BC which now will be seen as a C( j)-algebra.
Finally, if we begin with the D-valued quadratic form
z21 + z22
acting on the D-algebra D2 , then any of the two imaginary units, i or j, will arise
giving the factorizations into two factors each of which is D-two-dimensional.
k
z z0
e z0 z
z0
O 1 1 x
2
e†
−k
one-dimensional time with the x-axis and a one-dimensional space with the y-axis,
both axes being embedded into D and assume the speed of light to be equal to one;
then the set of zero-divisors with positive real part, x > 0 is nothing more than
the future of a ray of light which was sent from the origin towards either direction.
Analogously, the set of zero-divisors with negative real part, x < 0, represent the
past of the same ray of light.
Thus, positive hyperbolic numbers represent the future cone, i. e., they cor-
respond to the events which are in the future of the origin; the negative hyperbolic
numbers, that is, those which after multiplying by (−1) become positive, represent
the past cone, i.e., they correspond to the events in the past of the origin.
Accepting this interpretation, we see how the set of hyperbolic numbers which
are greater than a given hyperbolic number Z represents the events in the future
of the event which is represented by Z.
z ≺ y or z y or z = y.
z · y ∈ D+ .
z · y ≺ 0.
• If z ≺ y and w 0, then z · w ≺ y · w.
• If z ≺ y and w ≺ 0, then z · w y · w.
• If z is a (strictly) positive hyperbolic number, then it is invertible and its
inverse is also positive: if z 0 and z ≺ y, then y−1 0 and y−1 ≺ z−1 .
Example 2.6.2.1. Let us illustrate the above properties solving for z = β1 e + β2 e†
in D the inequality
|z|k w , (2.20)
44 Chapter 2. Algebraic Structures of the Set of Bicomplex Numbers
[a, b]D := z ∈ D a z b .
k z = β1 e + β2 e† 1
gives:
1 ≤ β1 ≤ 1, −1 ≤ β2 ≤ 1.
It turns out that in this case the hyperbolic interval is a one-dimensional set.
See the Figure 2.6.2.
• Take now a = k and b = 2 (obviously k ≺ 2). In this case the hyperbolic
interval is given by
O 1 1 x
2
e†
If A is D-bounded from below, then the inf D A can be computed by the formula
The above formulas explain a very peculiar character of the partial order on D.
Note only that although two D-upper (or D-lower) bounds can be incomparable
nevertheless they are always comparable with supD A (or with inf D A).
46 Chapter 2. Algebraic Structures of the Set of Bicomplex Numbers
k 2e
β1 e
−e† e
O 1 1 x
z = β1 e + β2 e†
2
e†
β2 e†
2e†
Let A and B be two subsets of D, denote by −A the set of all the elements
in A multiplied by −1, and denote by A + B the set of all sums z + y with z ∈ A
and y ∈ B; define in the same fashion the set A · B. The reader is invited to prove
the following properties.
• A set A is D-bounded from above (or from below) if and only if the set −A
is D-bounded from below (or from above); for such sets it holds that
• If A and B are D-bounded from below, then so is A + B and for such sets
one has:
inf D (A) + inf D (B) = inf D (A + B) .
If A and B are D-bounded from above, then so is A + B and for such sets
one has:
supD (A) + supD (B) = supD (A + B) .
|Z + W |k = |(β1 + ν1 ) · e + (β2 + ν2 ) · e† |k
= |β1 + ν1 | · e + |β2 + ν2 | · e†
(|β1 | + |ν1 |) · e + (|β2 | + |ν2 |) · e†
= |Z|k + |W |k .
The three properties (i)–(iii) manifest again the analogy between real positive
numbers and hyperbolic positive numbers; now we see that the hyperbolic modulus
of a bicomplex number has exactly the same properties as the real modulus of a
complex number whenever the partial order is used instead of ≤.
Because of properties (i)–(iii) we will say that | · |k is the hyperbolic-valued
(D-valued) norm on the BC-module BC.
It is instructive to compare (ii) with (1.13) where the norm of the product
and the product of the norms are related with an inequality. We believe that one
could say that the hyperbolic norm of bicomplex numbers is better suited to the
algebraic structure of the latter although, of course, one has to allow hyperbolic
values for the norm.
Remark 2.7.1. (1) Since for any Z ∈ BC it holds that
√
|Z|k 2 · |Z| , (2.21)
48 Chapter 2. Algebraic Structures of the Set of Bicomplex Numbers
In contrast with property (ii) above, this inequality involves both the Euclidean
and the hyperbolic norms.
(2) Take z1 and z2 in D+ , then clearly
(3) Note that the definition of hyperbolic norm for a bicomplex number Z does
not depend on the choice of its idempotent representation. We have used, for
Z ∈ BC, the idempotent representation Z = β1 e + β2 e† , with β1 and β2 in
C(i). If we had started with the idempotent representation Z = γ1 e + γ2 e† ,
with γ1 and γ2 in C( j), then we would have arrived at the same definition of
the hyperbolic norm since |β1 | = |γ1 | and |β2 | = |γ2 |.
(4) The comparison of the Euclidean norm |Z| and the D-valued norm |Z|k of a
bicomplex number Z gives:
1
||Z|k | = √ |β1 |2 + |β2 |2 = |Z| (2.23)
2
where the left-hand side is the Euclidean norm of a hyperbolic number.
• 1 ∈ D+
inv ;
• if λ ∈ D+
inv , then λ is invertible in D and λ
−1
∈ D+
inv .
+
It turns out that, anyway, both D+e and De† can be endowed with the struc-
ture of a multiplicative group. Beginning with D+ e one observes that for any λ > 0
1
one has that λe · e = λe, and e · λe = 1 · e = e. Hence, if we endow D+ e with the
λ
multiplication
which is the restriction of the hyperbolic multiplication, then on
(D+
e ,
) we have that
1. if λ1 e and λ2 e are in D+ +
e , then their product λ1 e
λ2 e = λ1 λ2 e is in (De ,
);
1 λe = e · λe = λe;
1
3. if λe ∈ (D+
e ,
), then e is its
-inverse:
λ
1
λe
e = 1 · e = e = 1 .
λ
D+
inv ⊂ Dinv ⊂ BCinv ,
where ⊂ means the embedding of group structures. At the same time we can use the
same arguments on De := e·D\{0} = {λe | λ ∈ R\{0} }, De† := e† ·D\{0} = {μe† |
μ ∈ R \ {0} }, BCe \ {0} := {λe | λ ∈ C \ {0} }, BCe† \ {0} := {μe† | μ ∈ C \ {0} }.
The first two of them are isomorphic to the multiplicative group of real numbers,
the last two are isomorphic to the multiplicative group of complex numbers.
We leave the details to the reader.
In addition to the references from the end of Chapter 1, such as [45, 56,
57, 58, 65, 85], where some of the algebraic structures that we developed in this
chapter have been introduced and studied, the most important contribution in
this area is the book of Alpay, Luna–Elizarrarás, Shapiro and Struppa [2]. This
study introduces for the first time the notion of the hyperbolic-valued norm on
the ring of bicomplex numbers, which we fully described in this chapter.
Chapter 3
The geometry of complex numbers coincides with the geometry of the Euclidean
space R2 , and this is because of a good compatibility between the algebraic struc-
ture of C and the geometry of R2 , which is expressed by the equality
z · z = x2 + y 2 = |z|2 . (3.1)
0 1 x
Using this segment, one constructs the two-dimensional cube pasting per-
pendicularly the cube of the dimension one at each point of [0, 1], i.e., pasting the
translations of [0, 1] and thus obtaining the square with unitary sides. See Figure
3.1.2.
···
0 1 0 1
0 0
1 1
The same idea allows us to visualize the whole 4-dimensional space with the
coordinate axes x0 , x1 , x2 , x3 : to each point of the axis x0 , we attach, perpendic-
ularly a copy of R3 . See Figure 3.1.5.
There is an alternative way of “seeing” the 4-dimensional world. Take, for
instance, the real 2-dimensional plane spanned by 1 and i; to each point of it we
attach perpendicularly a copy of the 2-dimensional plane spanned by j and k.
Note that whenever we mention the perpendicularity we mean, implicitly,
that the two sets pass through the origin and they are the orthogonal comple-
ments of each other; if they do not pass through the origin, then they become the
54 Chapter 3. Geometry and Trigonometric Representations
orthogonal complements of each other if they are translated to the origin. This
is reasonably clear on the level of intuition for the 3-dimensional situation, and
for the 4-dimensional case one can appeal to analogies in three dimensions; for
instance, an analogue of the two planes which are perpendicular in R4 may be two
straight lines which are perpendicular in R2 .
Traditionally, the 4-dimensional cube is presented as a solid in R4 whose 3-
dimensional surface consists of eight 3-dimensional cubes which are glued together
in a very particular way. We will see now that our approach does not contradict
this vision. The matter is that if instead of pasting a 3-dimensional cube at each
point of the interval [0, 1] on the axis x0 , the process can be repeated pasting cubes
at each point of the interval [0, 1] but now on the axis x1 , or on the axis x2 , or on
the axis x3 (we will denote for short the corresponding intervals as [0, i], [0, j] and
[0, k]), obtaining the same geometrical figure. With this in mind, let us introduce
the notations:
• Cr is the cube attached to the point r ∈ [0, 1].
• Is is the cube attached to the point s ∈ [0, i].
• Ju is the cube attached to the point u ∈ [0, j].
3.1. Drawing and thinking in R4 55
Proof. Follows by noting that the cubes C0 , etc., are the “extreme cubes” when
the described process applies to each of the following segments: [0, 1], [0, i], [0, j],
[0, k].
In order to see more precisely how the vertices of the cubes are glued together,
we will write down the vertices explicitly.
j+k
C0 i+j+k
1+j+k k
i+k
1+k j
0 i+j
i 1+i+j+k
1+j
1+i+j
1
1+i+k
C1 1+i
j+k
i+j+k
1+j+k
1+i+j+k
I0 k
j Ii
i+k
1+k i+j
1+j
0
1+i+j
i
1
i+1
1+i+k
i+k
j+k
Jj j+i+k
k
1+j+k 1+i+j+k
1+k
j
0 j+i
J0
i
j+1
1
1+i+j
1+i
1+i+k
k+j+1
k+j
k Kk
k+i k+i+j
j
1+i+j+k
0
k+1
i+j
1+j i
1+k+i K0
1
1+i+j
1+i
1 C1 I0 J0 K0
i C0 Ii J0 K0
j C0 I0 Jj K0
k C0 I0 J0 Kk
1+i C1 Ii J0 K0
1+j C1 I0 Jj K0
1+k C1 I0 J0 Kk
i+j C0 Ii Jj K0
i+k C0 Ii J0 Kk
j+k C0 I0 Jj Kk
1+i+j C1 Ii Jj K0
1+i+k C1 Ii J0 Kk
1+j+k C1 I0 Jj Kk
1+i+j+k C1 Ii Jj Kk
i+j+k C0 Ii Jj Kk
Since
2
2
z1 z2
+ = 1,
|Z|i |Z|i
the system
z1 z2
cos Θ = , sin Θ = , (3.6)
|Z|i |Z|i
has at least a solution Θ ∈ C(i). Actually there are infinitely many solutions, due
3.2. Trigonometric representation in complex terms 59
For the principal value of the complex argument, we will use the notation argC(i) (Z).
For a fixed m ∈ Z, we denote by argC(i),m (Z) := argC(i) (Z) + 2mπ, thus
ArgC(i) (Z) = argC(i),m (Z) m ∈ Z .
implies that r = |z| and α ∈ Arg(z). The bicomplex situation is more sophisticated.
If a bicomplex number has a representation of the form
Z = C · (cos α + j sin α)
with C and α complex numbers in C(i), then C is not necessarily the complex
modulus of Z; it is so if and only if C is in the upper half-plane and in this case
α ∈ ArgC(i) (Z). If C is in the lower half-plane, then |Z|i = −C and α + π ∈
ArgC(i) (Z): this is because
(a) if the complex number |Z|i · |W |i is in the upper half-plane, then indeed
|Z · W |i = |Z|i · |W |i ;
(b) if the complex number |Z|i · |W |i is not in the upper half-plane, then
|Z · W |i = −|Z|i · |W |i ;
Θ + Ω + π ∈ ArgC(i) (Z · W )
and
ArgC(i) (Z · W ) + π = ArgC(i) (Z) + ArgC(i) (W ).
but again not always |Z|ni and |Z n |i , as well as nΘ and ΘZ n , coincide. They do
coincide when the complex number |Z|ni is in the upper half-plane; if it is not,
then
|Z n |i = −|Z|ni and ΘZ n = nΘ + π.
We will call formula (3.9) the bicomplex De Moivre formula, although under-
standing that it does not always coincide with the trigonometric representation in
complex terms of Z n .
Since for an invertible Z there holds:
0 = Z · Z † = z12 + z22 ,
that is,
Z · Z † (z12 + z22 )−1 = Z · Z † |Z|−2
i = 1,
then using the trigonometric form of Z † we have:
Z −1 = Z † · |Z|−2
i = |Z|i (cos Θ − j sin Θ) · |Z|−2
i
62 Chapter 3. Geometry and Trigonometric Representations
and finally
Z −1 = |Z|−1
i (cos Θ − j sin Θ) . (3.10)
With this, formula (3.9) becomes true for any integer number n.
This result allows us to write the quotient of two invertible bicomplex num-
bers in trigonometric form:
Z |Z|i
= (cos(Θ − Ω) + j sin(Θ − Ω)) . (3.11)
W |W |i
Z
= cos(2Θ) + j sin(2Θ) , (3.12)
Z†
which is, obviously, a bicomplex number of complex modulus 1.
An analogous trigonometric form is obtained if we work with |Z|j . The defi-
nitions and computations are completely similar, yielding the trigonometric form
Z = β 1 e + β 2 e†
which leads to
β1 β2 †
Z = |Z|k · ·e+ ·e , (3.14)
|β1 | |β2 |
β1 β2
where the C(i)-complex numbers and are of (real) modulus one. Thus,
|β1 | |β2 |
they are of the form
β1 β2
= eiθ1 , = eiθ2 (3.15)
|β1 | |β2 |
{θ1 , θ2 } ⊂ [0, 2π), hence θ1 and θ2 have a well-defined and well-known geometric
meaning. Formula (3.14) leads us to consider the hyperbolic number
ΨZ := θ1 e + θ2 e†
Ψ0 := ν1 e + ν2 e†
such that
z1 z2
cos Ψ0 = and sin Ψ0 = ; (3.17)
|Z|k |Z|k
64 Chapter 3. Geometry and Trigonometric Representations
with ΨZ = ν1 · e + ν2 · e† ∈ D+
inv being the hyperbolic angle, or hyperbolic principal
argument, of Z.
Proof. It is known (see again Chapter 6 but also [45]) that
and
sin Ψ0 = sin(ν1 e + ν2 e† ) = sin ν1 · e + sin ν2 · e† ;
hence
Taking into account the periodicity of trigonometric functions and choosing the va-
lues of ν1 and ν2 in the interval [0, 2π) we obtain that Ψ0 may be taken equal to ΨZ .
We prefer the notation ΨZ emphasizing that the angle is linked to Z ∈ BC\S0 .
In contrast with the complex argument of a bicomplex number which is de-
fined for invertible bicomplex numbers only, we can define the hyperbolic argument
for zero-divisors too. Consider, for instance, a zero-divisor Z = β1 e, then
Z · Z ∗ = |β1 |2 e
β1
Z = |Z|k · · e,
|β1 |
3.3. Trigonometric representation in hyperbolic terms 65
β1
and since is a complex number of (real) modulus one, it can be written as
|β1 |
iν1
e with ν1 ∈ [0, 2π). In this case the hyperbolic angle associated to Z is the
hyperbolic zero-divisor
ν := ν1 e.
Analogously, if Z = β2 e† , then |Z|k = |β2 |e† and
β2
Z = |Z|k · · e† = |Z|k eiν2 e† .
|β2 |
In this case, the hyperbolic angle associated to Z is the hyperbolic zero-divisor
ν = ν 2 e†
for an invertible Z and similarly for a zero-divisor. Finally, ArgD Z denotes the set
of all possible hyperbolic angles:
ArgD Z := argm,n;D Z m, n in Z
thus, the usual trigonometric form of the complex number z and its trigono-
metric form in hyperbolic terms when seen as a bicomplex number, coincide.
In particular, the hyperbolic argument of z, Ψz = θe + θe† = θ, coincides
with the usual (real) argument of z. Hence
(b) The imaginary unit j has the idempotent form j = (−i)e + ie† , thus |j|k =
1e + 1e† = 1. Hence, its trigonometric representation in hyperbolic terms is
3 π
j = |j|k −ie + ie† = ei 2 π e + ei 2 e† ,
3 π
and its hyperbolic argument is Ψj = πe + e† , thus
2 2
3 1
ArgD j = + 2m πe + + 2n πe† m, n ∈ Z .
2 2
(c) If Z = z = ae + be† , with a, b in R, that is, z ∈ D, |z|k = |a|e + |b|e† , then its
trigonometric representation in hyperbolic form is
a b †
z = |z|k e + e = |z|k ±e ± e† .
|a| |b|
Hence there are four options for the principal value of the hyperbolic argu-
ment of a hyperbolic number as indicated in Figure 3.3.1:
Ψz = 0e + 0e† = 0,
or
Ψz = 0e + πe† = πe† ,
or
Ψz = πe + 0e† = πe,
or
Ψz = πe + πe† = π.
Hence the set ArgD z is, according to the quadrant the hyperbolic number
z is in, as follows:
ArgD z
⎧
⎪
⎪ 2π(me + ne† ) m, n ∈ Z if z is in the 1st quadrant;
⎪
⎪
⎪
⎨ 2πme + (2n + 1)πe m, n ∈ Z
†
if z is in the 2nd quadrant;
=
⎪
⎪ (2m + 1)πe + 2nπe† m, n ∈ Z if z is in the 3rd quadrant;
⎪
⎪
⎪
⎩
2nd
Ψz = πe†
3rd e 1st
Ψz = π Ψz = 0
O x
e†
Ψz = πe
4th
Similarly for the case when one of the factors, or both, are zero-divisors.
In particular, taking Z = W one gets that
Z 2 = |Z|2k · ei(2ν1 ) · e + ei(2ν2 ) · e†
= |Z|2k · (cos(2ν1 ) + i sin(2ν1 )e + (cos(2ν2 ) + i sin(2ν2 )e† .
which implies immediately that the De Moivre formula is true for any n ∈ Z.
Let us see what is the relation between the trigonometric representations
in hyperbolic terms of Z and Z ∗ . Since the ∗-conjugate of Z = β1 e + β2 e† is
Z ∗ = β 1 e + β 2 e† , we have:
thus
β1 β
Z ∗ = |Z ∗ |k · · e + 2 · e†
|β1 | |β2 |
−iν1 −iν2
= |Z|k · e ·e+e · e† ,
one can formally define the bicomplex sphere Sγ0 centered at the origin and with
radius being a fixed positive hyperbolic number γ0 = a0 e + b0 e† ∈ D+ , i.e.,
Sγ0 := Z ∈ BC |Z|k = γ0 .
this set is a circumference in the real two-dimensional plane BCe with center at
a0
the origin and radius √ = |a0 · e|. Similarly, if γ0 = b0 · e† , then the set Sγ0 is a
2
b0
circumference in BCe† with center at the origin and radius √ = |b0 · e† |.
2
Thus, whenever the radius of a bicomplex sphere is a (positive) zero-divisor,
i.e., a semi-positive hyperbolic number, then the sphere is, in fact, a usual circum-
ference. See Figures 3.3.2 and 3.3.3.
C(i)
i
ie a0 e
e
O 1 a0
O BCe
C(i)
i ie†
e† b 0 e†
O 1 b0 O
BCe†
a0
√
2
b0
√
2
corresponding circumference:
iμ1 a0
a0 e e ∈ Ce 0; √ ,
2
b0
b0 eiμ2 e† ∈ Ce† 0; √ ;
2
and moreover,
a0 eiμ1 e + b0 eiμ2 e† = a0 e + b0 e† · eiμ1 e + eiμ2 e†
= γ0 · eiμ1 e + eiμ2 e† = Z.
This shows that the value of the hyperbolic modulus of Z tells us on which sphere
Z is situated and its hyperbolic argument determines the exact place of Z on the
sphere.
Note also that the sphere Sγ0 is in a bijective correspondence with the hy-
perbolic interval
[0, 2π)D = z ∈ D 0 z ≺ 2π
= z = ν1 e + ν2 e† ν1 , ν2 ∈ [0, 2π) ⊂ D+
2πe
D
2π
2πe†
a0 eiν1 e + b0 eiν2 e† ←→ ν1 e + ν2 e† .
We will see later that this is related with the exponential functions of bicomplex
and hyperbolic variables.
Some of the geometric and trigonometric properties of hyperbolic and bicom-
plex numbers have been studied in the literature in works such as [28, 45, 56, 100].
We also mention here [87] and [11], where trigonometric representations of hy-
perbolic and hypercomplex numbers are studied, together with applications to
Special Relativity. The authors of [11] emphasize that the hyperbolic numbers are
“the mathematics of the two-dimensional Special Relativity”, and adapting this
two-dimensional hyperbolic geometry to “multidimensional commutative hyper-
complex systems” (such as bicomplex numbers) leads to a concrete application to
Special Relativity in four-dimensional Minkowski space-time. The “separation” of
the four-dimensional bicomplex space into spacelike and timelike parts (events)
is captured by the geometry of the complex lines Le and Le† , which we carefully
study in complete detail in the next chapter.
Chapter 4
a 1 x + a2 y = b , (4.1)
where a1 , a2 , b are real coefficients. Writing a = (a1 , a2 ), z = (x, y), (4.1) is equiv-
alent to
a, z (a1 , a2 ), (x, y)R2 b
= = . (4.2)
| a | | a | | a |
This formula has the following geometric description: a point (x, y) belongs to
the straight line given by (4.1) if the corresponding vector has a constant projection
b
on the straight line determined by a.
|a|
1
Setting z := x + iy and a := a1 + ia2 and recalling that x = (z + z),
2
i 1 i
y = (z − z), a1 = (a + a) and a2 = (a − a), we obtain the complex form of
2 2 2
writing equation (4.1):
az + az = 2b . (4.3)
Geometrically, a complex number z belongs to the line given by (4.3) if
Re(az) = Re(az) is equal to the constant b (see Figure 4.1.1).
āz
āz + az̄ = 2b
z
ā
2b b
a
z̄
az̄
a 2b
z=− z+ . (4.6)
a a
We introduce the notation
a
M := − = −ei2φ0 = ei(2φ0 +π) .
a
a2
Since a = |a|2 a−1 , then M = − . Moreover, in order to capture the slope
|a|2
of our line, we write:
z =Mz+B, (4.7)
2b
where B := .
a
We note here some particular cases. If M = −1, then a = a = a1 , a real
number, i.e., a2 = 0, which yields the vertical line a1 x = b, which has no well-
defined slope. In complex terms the equation of this vertical line is z = −z + B,
2b
where B = is a real number.
a1
If M = 1, then a = −a = ia2 , so a1 = 0, which yields the horizontal line
2b
a2 y = b. In complex terms we obtain z = z + B, where B = i is a purely
a2
imaginary number.
If it is known that the real line passes through a point z0 = x0 + iy0 and it
has slope m = 0, then the equation can be written as
y − y0 = m(x − x0 ) ,
or in complex terms as
z + z0 = M (z + z0 ) .
If the real line is determined by two points z0 = x0 + iy0 and z1 = x1 + iy1 ,
such that x0 = x1 , it is given by equation
y1 − y0
y − y0 = (x − x0 ) . (4.8)
x1 − x0
76 Chapter 4. Lines and curves in BC
z0
ia
a z0 + tia
ia
tia
Following the analogy with the real lines, we define a complex straight line
in C2 (i) to be the set of solutions (z1 , z2 ) ∈ C2 (i) of the equation
a1 z1 + a2 z2 = b, (4.11)
A† Z + AZ † = 2b, (4.12)
0 · z 1 + 1 · z2 = 0 ,
1 · z1 + 0 · z2 = 0 ,
and let us show that it is a C(i)-complex line which we will call the C(i)-complex
line generated by Z0 and passing through the origin. Indeed, for any element
Z = z1 + jz2 in LZ0 there exists λ ∈ C(i) such that
z2 z2
Assuming that z20 = 0, one has that λ = 0 and hence z1 = 0 z10 , or
z2 z2
equivalently:
z1 z20 − z2 z10 = 0 .
This equation is a particular case of (4.11) with a1 = z20 , a2 = −z10 and b = 0.
If z20 = 0, the set
Lz1 := {λz10 λ ∈ C(i)} = C(i)
z1 a 1 + z 2 a 2 = 0 (4.13)
with a1 = 0, then one can take Z0 := −a2 + ja1 ; the set of solutions Z = z1 + jz2
of (4.13) coincides with LZ0 .
Example 4.1.4.1. We know already that any zero-divisor is of the form β1 e or
β2 e† , that is, Z is a zero-divisor if and only if Z ∈ Le \ {0} or Z ∈ Le† \ {0}, thus
the set of zero-divisors (together with the zero) is described by
S0 = Le ∪ Le† .
Note also that here we have two complex lines (Le and Le† ) that intersect only at
the origin. The reader should recall that in R3 a pair of two-dimensional planes
cannot intersect at just one point, but in BC ∼= R4 there is an additional dimension
which allows this phenomenon.
The corresponding equations of the complex lines Le and Le† are, respec-
tively,
z1 + iz2 = 0
and
z1 − iz2 = 0 .
In order to define a complex line that does not necessarily pass through the
origin, take W0 = w10 +jw20 ∈ BC\{0}. Then the C(i)-complex line passing through
W0 and parallel to LZ0 (that is, a complex line with the direction determined by
Z0 ) is defined as the set
LZ0 + W0 := {λZ0 + W0 λ ∈ C(i)} .
80 Chapter 4. Lines and curves in BC
Let us show that this definition is equivalent to that of a complex line. Indeed,
given a point Z ∈ LZ0 + W0 , with Z = z1 + jz2 , there exists λ ∈ C(i) such that
z1 = λz10 + w10 ,
z2 = λz20 + w20 .
Assuming again that z20 = 0, we obtain that the complex cartesian components of
Z satisfy the equation
z1 a1 + z2 a2 = b, (4.14)
with a1 = 0 and a2 = 0, let us find Z0 and W0 in BC\{0} such that any Z = z1 +jz2
whose components satisfy (4.14) belongs to LZ0 + W0 . From the above equation
we can write z1 as
b − a2 1 − z2
z1 = + a2 .
a1 a1
z2 − 1 b − a2
Define λ := , that is, z1 = − λa2 . Substituting in (4.14), we write
a1 a1
now z2 as
z2 = 1 + λa1 .
b − a2
Let us show that Z0 := −a2 +ja1 and W0 := +j are the bicomplex numbers
a1
which we are looking for; indeed,
b − a2
λZ0 + W0 = −λa2 + jλa1 + +j
a
1
b − a2
= − λa2 + j (λa1 + 1)
a1
= z1 + jz2 = Z,
Lj + j with Lj := λj λ ∈ C(i) .
a1
Let us come back to equation (4.12) where the coefficient A is defined to be
A = a1 + ja2 ; we know already that the bicomplex number Z0 which determines
the direction of the complex line is given as
Z0 = −a2 + ja1 = j(a1 + ja2 ) = jA ,
that is,
A = −jZ0 .
Thus (4.12) becomes
Z0 Z † − Z0† Z = 2jb . (4.16)
If we perform a similar analysis in C2 (j), then all the formulas and conclusions
above are valid in their corresponding analogues. For example, the C( j)-complex
line
α1 ζ1 + α2 ζ2 = ν,
where all numbers and coefficients are in C(j), has the bicomplex formulation
AZ + AZ = 2ν,
where A = α1 + iα2 and Z = ζ1 + iζ2 are bicomplex numbers. Nevertheless, they
are in the same set BC, so that, as a matter of fact, there are two types of complex
lines in BC, the C(i)-lines and C(j)-lines.
Given a C(i)-complex straight line, is it possible to consider it as a C(j)-
complex straight line or reciprocally? This question is related with the following
one: given a real two-dimensional plane in BC, is it always possible to consider it
as a complex line?
Here one can find some partial answers to these questions:
• What are the C(i)-complex lines containing the number 1? Assume that L is
such a C(i)-complex line of the form L = LZ0 for some Z0 ∈ BC \ {0}. Let
L1 be the complex line generated by 1, then obviously L1 = C(i). We know
also that 1 ∈ LZ0 is equivalent to LZ0 = L1 , that is, L = C(i).
Thus, among all the two-dimensional real planes which contain the axis x1 ,
there is only one which has the structure of a C(i)-complex line.
• In the same way, among all the two-dimensional real planes which contain
the axis x1 , only C(j) is a C(j)-complex line.
• Other conclusions are: C(j) is not a C(i)-complex line; C(i) is not a C(j)-
complex line.
• As we did with the C(i)-complex straight lines, it can be proved in a similar
fashion that, given Z0 ∈ BC \ {0}, the C(j)-complex line generated by Z0 ,
i.e., in the direction of Z0 , is the set
TZ0 := {μZ0 μ ∈ C(j)} .
where Θ and Θ0 are the complex arguments of A and Z, respectively. Then equa-
tion (4.12) becomes
|A|i |Z|i cos(Θ − Θ0 ) = b. (4.18)
We observe here the similarity between the equations of real lines (4.4) and
complex lines (4.18)! Moreover, inspired by (4.18), we define the C(i)-complex
projection of Z onto A, for any invertible bicomplex numbers A and Z, by
where Θ − Θ0 can be called the complex angle between A and Z. We will come
back to this notion later.
LZ0 + W0 = C(i) + W0 ,
b
with W0 = j. This is a complex line which we will call parallel to the complex
a2
line C(i). Making the convention that the complex line C(i) determines the notion
of C(i)-complex horizontality, any complex line parallel to C(i) will be called C(i)-
complex horizontal.
It follows that |A|i = ±a2 , where the sign should be chosen in accordance
with our agreement that the imaginary part of the complex number |A|i is non-
π
negative. In this case Θ0 = becomes a real number.
2
84 Chapter 4. Lines and curves in BC
Notice that for any μ = a + ib ∈ C(i) the product μj is μj = aj + kb, hence the
C(i)-complex line Lj is the two-dimensional real plane generated by j and k. In
this case, our complex line is of the form
L Z0 + W 0 = L j + W 0
b
with W0 = + j.
a1
Note that for the set Lj + W0 one has:
b
Lj + W0 = {μ · j + + j μ ∈ C(i)}
a1
b b
= {(μ + 1) · j + μ ∈ C(i)} = {μ · j + μ ∈ C(i)}
a1 a1
b
= Lj + .
a1
b b
In other words, one may take W0 = and thus any point Z ∈ Lj + is of the
a1 a1
b b
form Z = μ · j + , i.e., Z = z1 + jz2 with a constant z1 = and an arbitrary
a1 a1
complex number z2 .
Making now the convention that the complex line Lj determines the notion
of C(i)-complex verticality, any C(i)-complex line Lj + W0 , that is parallel to Lj ,
will be called C(i)-vertical.
Since in this case |A|i = ±a1 and Θ0 can be taken to be zero, then the
equation of any vertical line in trigonometric form is
Since Lj ∩ C(i) = {0}, then any horizontal line intersects any vertical line in just
one point. We have seen this phenomenon before, in the sense that R4 is “wide”
enough and a pair of two-dimensional planes may intersect at one point only.
Now some words about the case when Z0 is a zero-divisor. By Proposi-
tion 4.1.5.2, this is equivalent to LZ0 = Le or LZ0 = Le† , thus, if Z0 ∈ S,
then any C(i)-complex line LZ0 + W0 is parallel to Le or to Le† . We call these
types of complex lines zero-divisor complex lines.
4.1. Straight lines in BC 85
We will see now that it is quite appropriate to say that a complex line L :=
LZ0 + W0 has the direction along Z0 (or that it is parallel to LZ0 ) since we will
relate the complex line with the complex argument of Z0 .
Recall that Z0 = −a2 + ja1 , with
a 1 z1 + a 2 z2 = b (4.19)
being the equation of a complex line parallel to LZ0 and passing through W0 . Let
us start with the case Z0 ∈ S0 , which is equivalent to Z0 Z0† = 0; thus, we can
consider the trigonometric form of Z0 :
a2 a1
Z0 = |Z0 |i − +j = |Z0 |i (cos Ψ0 + j sin Ψ0 ) .
|Z0 |i |Z0 |i
which we will call the complex slope of the line L. Comparing with equation (4.19),
we have a deep analogy with the definition of slope of usual real lines in R2 . Thus,
the slope Λ of the complex line L is the tangent of the complex argument Ψ0 of
Z0 , with Z0 ∈ S0 being a bicomplex vector that determines the direction of the
line L.
Note that if Z0 = −a2 + ja1 is still a non-zero-divisor, but a2 = 0, we have
seen before that in this case we have a vertical line, and again the analogy with real
lines in R2 is maintained, since we can define Λ = ∞ and the complex argument
π
of Z0 is , coinciding with the usual notion of a vertical line.
2
What happens when Z0 is a zero-divisor? We know already that LZ0 = Le
or LZ0 = Le† and our line is a zero-divisor complex line. Since in this case we have
also that
1 1 1
Z0 = −a2 + ja1 = λ (1 ± ij) = λ ± iλ · j, λ ∈ C(i),
2 2 2
then
a1
Λ=− = ±i
a2
and since there are no complex numbers Ψ such that tan Ψ = ±i we can define
the slope of the zero-divisor complex lines as being equal to i or −i.
The previous reasoning motivates the question about the relationship be-
tween complex arguments of bicomplex numbers belonging to the same complex
line.
86 Chapter 4. Lines and curves in BC
In other words, the slope of LZ0 does not depend on the generator Z0 .
Proof. Given S ∈ LZ0 \ {0}, then there exists λ ∈ C(i) \ {0} such that
S = λZ0 = λ z10 + jz20 = λz10 + jλz20 ;
hence S ∈/ S0 . We know already that |S|i = ±|λ|i · |Z0 |i . More precisely we have
to consider the following cases.
(a) If λ ∈ Π+ , the
upper
half-plane,
this is equivalent to |λ|i = λ, and thus
ArgC(i) (λ) = 2πm m ∈ Z .
(a.1) If |λ|i · |Z0 |i = λ|Z0 |i ∈ Π+ , then |S|i = λ|Z0 |i and thus
We are able now to provide different ways of writing the equation of a complex
line, according to the information that one has at hand. Consider first the case in
which L is not a vertical line, thus a2 = 0, and write the equation of L as
a1 b
z2 = − z1 + = Λz1 + c , (4.20)
a2 a2
b
with c = . Let Z = z1 + jz2 be such that its cartesian components satisfy (4.20).
a2
Using also the relations
1 j †
z1 = (Z + Z † ), z2 = (Z − Z) , (4.21)
2 2
one obtains:
Z = ρZ † + C , (4.22)
Λ−j 2b
with ρ := − and C := .
Λ+j a2 (Λ + j)
Assume now that we have a line L with slope Λ = Λ1 + jΛ2 = 0 and passing
through the point S0 = s01 + js02 with s01 and s02 in C(i). We know that L is of the
form LZ0 + S0 . Note also that if Z = z1 + jz2 ∈ L, then Z − S0 ∈ LZ0 ; since
z2 − s02
Λ= ,
z1 − s01
Using again (4.21) and the recently defined ρ, the previous equation can be written
in terms of Z and Z † :
Z − S0 = ρ(Z † − S0† ) . (4.24)
Here one notes that if L passes through the origin, one may take S0 = 0, and
any non-zero Z ∈ L satisfies then
Z Λ−j
†
=ρ=− .
Z Λ+j
Following what one does for real lines in R2 , we would like now to find the
equation of a complex line passing through two given points. We need first a few
results.
88 Chapter 4. Lines and curves in BC
iS = (iλ)Z0 ∈ LZ0 = L .
For the reciprocal statement, given any Z0 ∈ L, the set {Z0 , iZ0 } is a linearly
independent set in R4 , thus the R-linear subspace L{Z0 , iZ0 } generated by it
coincides with L. Thus L = LZ0 is a complex line.
Proposition 4.1.7.4. If there are two points Z and W in a real two-dimensional
plane L containing the origin, such that the difference of their complex arguments
is neither zero nor π, then L is not a C(i)-complex line.
Proof. Follows directly from Proposition 4.1.7.1.
Proposition 4.1.7.5. Given Z0 ∈ BC \ {0}, there exists a unique C(i)-complex line
L passing through Z0 and the origin.
Proof. The proof follows from the fact that L is a real two-dimensional plane
containing Z0 and iZ0 .
Let us come back to the task of finding the equation of the complex line
passing through two given points S and W . Set Z0 := S − W and let LZ0 be the
unique C(i)-complex line that contains Z0 . If S = s01 + js02 and W = w10 + jw20 ,
s0 − w20
then Z0 = (s01 − w10 ) + j(s02 − w20 ), thus the slope of LZ0 is Λ = 20 . Having
s1 − w10
the slope and the point S that belongs to L, and using (4.23), the equation of L
is:
s0 − w20
z2 − s02 = 20 · (z1 − s02 ) . (4.25)
s1 − w10
with the bicomplex number Z = x1 + iy1 + jx2 + ky2 = (x1 + ky2 ) + i(y1 − kx2 ) =
z1 + iz2 ∈ BC, thus we will say also that Z is a solution of (4.26).
Since Z ∗ = z1 − iz2 , hence
1 i
z1 = (Z + Z ∗ ) and z2 = (Z ∗ − Z)
2 2
and (4.26) becomes
A∗ Z + AZ ∗ = 2b, (4.27)
with the bicomplex coefficient A := a1 + ia2 .
Consider as an illustration some particular cases.
1) If a1 = 0, a2 = 1, b = 0, then z2 = 0 and z1 takes any hyperbolic value, that
is, the set of solutions of the equation
z2 = 0
is the set z1 z1 ∈ D = D.
2) If a1 = 1, a2 = 0, b = 0, we have the equation
z1 = 0,
and
Z = z1 + iz2 = (η11 + iη12 )e + (η21 + iη22 )e† .
We observe that the component (η11 + iη12 )e = η11 e + η12 ie can be seen as a point (or
vector) in the real plane BCe (which is a complex line) with coordinates (η11 , η12 );
the same comment applies to the other component of Z as well as to both com-
ponents of A.
90 Chapter 4. Lines and curves in BC
PZ0 := μZ0 μ ∈ D
a1 z1 + a2 z2 = 0, (4.29)
or, equivalently,
m1 (u01 + iv10 ) = u1 + iv1 , m1 ∈ R,
m2 (u02 + iv20 ) = u2 + iv2 , m2 ∈ R;
this means that the pairs (u1 , v1 ) and (u2 , v2 ) are solutions, respectively, of the
equations
u01 u1 + (−u01 )v1 = 0,
(4.31)
u02 u2 + (−u02 )v2 = 0.
Hence Z belongs to PZ0 if and only if its components satisfy the equations (4.31).
92 Chapter 4. Lines and curves in BC
Comparing systems (4.31) and (4.28), we see that (4.31) determines a hyper-
bolic line if and only if it satisfies Corollary 4.2.2.
On the other hand the set of solutions of (4.31) admits descriptions in hy-
perbolic and bicomplex terms as in equations (4.26) and (4.27). To see this we
compute:
Z0 = z1 + iz2 = (u01 + iv10 )e + (u02 + iv20 )e†
= i(v10 + i(−u01 ))e + i(v20 + i(−u02 ))e†
= i(v10 e + v20 e† ) − (−u01 e − u02 e† )
=: ia1 − a2 = i(a1 + ia2 ) =: iA,
from where a1 = z02 and a2 = −z01 . Thus, the set PZ0 is the set of solutions of the
equation
z02 z1 − z01 z2 = 0.
Finally we know that such a set determines a hyperbolic line if and only if z02 and z01
satisfy the restrictions (i)–(vi) and this is equivalent to saying that Z0 ∈
/ S0 .
PZ0 + W0 := μZ0 + W0 μ ∈ D .
We leave it to the reader to show that any hyperbolic line can be described as
PZ0 + W0 for some bicomplex numbers Z0 , W0 .
We also leave it to the reader to prove the following results, which are anal-
ogous to those given in the case of complex lines.
Proposition 4.2.2.1. Given the hyperbolic line PZ0 and given U ∈ PZ0 such that
U = μ0 Z0 , with μ0 ∈ / S0 (hence U ∈ / S0 ), then PU = PZ0 . Reciprocally, if
PU = PZ0 , then U ∈ PZ0 .
Another characterization of real two-dimensional planes passing through the
origin and being a hyperbolic line is given in the following Proposition whose proof
is left to the reader.
Proposition 4.2.2.2. A real two-dimensional plane T ⊂ BC with 0 ∈ T is a hyper-
bolic line if and only if U ∈ T implies kU ∈ T .
Corollary 4.2.2.3. A hyperbolic line P through the origin, regarded as a two-
dimensional subspace of R4 ∼ = BC, is generated over R by any U0 ∈ P \ S0 and by
kU0 , i.e., it is the real span of U0 and kU0 .
Now, in analogy with complex lines, we describe the relation between the
hyperbolic arguments of the points of a hyperbolic line PZ0 .
4.2. Hyperbolic lines in BC 93
Theorem 4.2.2.4. Consider the hyperbolic line PZ0 for a point Z0 ∈ BC\S0 . Write
0 0
Z0 = |Z0 |k eiν1 e + eiν2 e† ,
We are now ready to define the slope of hyperbolic lines; Theorem 4.2.2.4
ensures that the slope will not depend on the bicomplex number Z0 that determines
its direction.
We now need to use a result about trigonometric functions of a hyperbolic
variable, a topic that is studied in detail in Chapter 6. As we will show there, the
tangent of a hyperbolic number ν = ν1 e + ν2 e† can be given a rigorous definition
and it turns out that
tan ν = tan ν1 e + tan ν2 e† .
(b) Each real line L in C(i) is exactly the intersection of a hyperbolic line P and
C(i) itself, and the slope of the hyperbolic line P coincides with the (usual
real) slope of L. Indeed, assume that 0 ∈ L, then any z ∈ L is of the form
z = |z| (cos θ + i sin θ) = |z|k eiθ e + eiθ e† ;
(c) Consider the hyperbolic line of the type Pie+β20 e† or Pβ10 e+ie† ; since the hy-
π π
perbolic arguments are Ψie+β20 e† = e + ν20 e† and Ψβ10 e+ie† = ν10 e + e† ,
2 2
then the slopes are not well defined but can be symbolically represented as
We say in this case that these hyperbolic lines determine the notion of hy-
perbolic verticality.
Recalling that a hyperbolic line is a real two-dimensional plane such that
its projections onto BCe and BCe† are usual real lines, we can give a precise
geometrical notion of the hyperbolic angle between two hyperbolic lines. Clearly
it is enough to define the hyperbolic angle between hyperbolic lines that pass
through the origin.
Definition 4.2.2.7. Take hyperbolic lines PZ0 and PW0 ; if argD Z0 = ν10 e + ν20 e†
and argD W0 = μ01 e + μ02 e† , then the (trigonometric) hyperbolic angle α between
PZ0 and PW0 is
that determines the orientation of the hyperbolic angle. The reason for this is
that in the projections on BCe and BCe† the roles of L1Z0 , L1W0 and L2Z0 , L2W0
do not coincide in general; just recall that D is not a totally ordered set but a
partially ordered one. That is why we use not the difference of the arguments but
its hyperbolic modulus.
with φ1 : [a, b] → C(i) and φ2 : [c, d] → C(i) being parametrizations of usual curves
in C(i).
Γ = γ 1 e + γ 2 e† .
96 Chapter 4. Lines and curves in BC
γ 1 e + γ 2 e†
γ1 BCe
γ2
BCe†
and r ∈ {1, 2, . . . t}; note that their union is a set of measure zero (as a subset of
the manifold Γ).
Of course, one may define more exotic types of hyperbolic curves eliminating
other restrictions that we imposed on γ1 and γ2 , but for our purposes the definition
given above is sufficient.
γ 0
β 2 e† S
T
0
β1 e
0 0
Z 0 = β 1 e + β 2 e†
W = T e + Se†
BCe S2
T2
λ1
S1 T1
θ1 θ2
γ1
β10 β20
BC†e
γ2
λ2
Figure 4.3.3: The hyperbolic angle between hyperbolic curves.
ϕ(u, v) = ψ1 (u, v) + jψ2 (u, v) := (ϕ1 (u, v) + iϕ2 (u, v)) + j (ϕ3 (u, v) + iϕ4 (u, v)) ,
with a plane T1 passing through the origin. Thus, any point W ∈ T can be written
as
∂ϕ1 ∂ϕ1 ∂ϕ2 ∂ϕ2
W = Z0 + a (Z0 ) + b (Z0 ), a (Z0 ) + b (Z0 ),
∂u ∂v ∂u ∂v
(4.34)
∂ϕ3 ∂ϕ3 ∂ϕ4 ∂ϕ4
a (Z0 ) + b (Z0 ), a (Z0 ) + b (Z0 ) .
∂u ∂v ∂u ∂v
Recall (see Section 4.1.3) that T is a complex line if and only if T1 is a complex
line passing through the origin and that T1 is a complex line if and only if it is
invariant under the multiplication by i, that is, T1 as a subspace of dimension two
100 Chapter 4. Lines and curves in BC
and iZ = −y1 + ix1 − jy2 + kx2 , i.e., the vector in (4.35) corresponds to iZ.
Thus, applying Mi to a point w1 ∈ T1 we get:
∂ϕ2 ∂ϕ2 ∂ϕ1 ∂ϕ1
Mi w 1 = −a (Z0 ) − b (Z0 ), a (Z0 ) + b (Z0 ),
∂u ∂v ∂u ∂v
(4.36)
∂ϕ4 ∂ϕ4 ∂ϕ3 ∂ϕ3
−a (Z0 ) − b (Z0 ), a (Z0 ) + b (Z0 ) .
∂u ∂v ∂u ∂v
Sγ0 := {Z ∈ BC | |Z|k = γ0 }
a0
origin and radius √ , contained in the complex line BCe . Similarly, if γ0 = b0 e† ,
2
b0
the sphere Sb0 e† is the circumference with center at the origin and radius √ ,
2
contained in the complex line BCe† . Finally, if γ0 is not a zero-divisor, recall that
the sphere Sa0 e+b0 e† is the surface of a torus:
i.e., if we are looking at BC as C2 (i) with the idempotent coordinates, then the
bicomplex ball Bγ0 is the bicomplex form of writing for the usual bidisk centered
at the origin and with bi-radius (a0 , b0 ).
If γ0 is a zero-divisor, then we cannot define the ball in the same way because
none of the inequalities |β1 | < 0 or |β2 | < 0 has solutions. So we define in this case
the ball Bγ0 to be one of the two disks: one is located in BCe with center at the
a0
origin and radius √ , and the other is located in BCe† with center at the origin
2
b0
and radius √ .
2
It is worth noting that for a bicomplex ball Bγ0 , with γ0 not a zero-divisor,
the respective bicomplex sphere Sγ0 is not its topological boundary but it is its
distinguished, or Shilov, boundary.
Sλ := Z = β1 e + β2 e† | |Z|k = λ
|Z · W |k = |Z|k · |W |k (4.37)
1) 1 ∈ S1 ;
2) if Z ∈ S1 , then it is invertible and Z −1 ∈ S1 ;
moreover, formula (4.37) says that
3) if Z, W ∈ S1 , then Z · W ∈ S1 .
Thus, the bicomplex unitary sphere S1 is a multiplicative group, but this is
an exceptional case for bicomplex spheres having a non-zero-divisor as its radius,
and for other non-zero-divisor values of the parameter λ the sphere Sλ is not a
multiplicative group; the same as in R2 .
Recall what happens when λ is a positive zero-divisor, i.e., when λ = λ1 e ∈
+
D+ † +
e , with λ1 > 0, or when λ = λ2 e ∈ De† , with λ2 > 0. If λ ∈ De , then for the
†
points Z = β1 e + β2 e of Sλ we have that
that is,
|β1 | = λ1 , |β2 | = 0,
thus, the bicomplex sphere Sλ is
Sλ = {Z | Z = β1 e, |β1 | = λ1 > 0 },
and it can be seen as the circumference of radius λ1 centered at the origin which is
located in the real two-dimensional plane, spanned by e and ie (which is a complex
line in BC).
Take λ1 = 1 here, then Sλ becomes
Se = {Z | Z = β1 e, |β1 | = 1 } .
We are now in a situation very similar to that of Section 2.7.1, and we will use a
similar approach in order to make Se a multiplicative group. The set Se is endowed
with the multiplication which is the restriction of the bicomplex multiplication
and which acts invariantly on Se ; then the element e is the neutral element for ,
1
every element β1 e in Se is -invertible and its -inverse is e ∈ Se .
β1
Thus, we conclude that (Se , ) is a multiplicative group which is not a sub-
group of the multiplicative group (BC \ S0 , ·). Of course, (Se , ) is isomorphic to
the multiplicative group of complex numbers of modulus one.
Note that if λ1 = 1, then Sλ1 e is not a multiplicative group. In other words,
among all the spheres Sλ1 e with λ1 > 0, only one, Se , is a multiplicative group.
The situation with λ = λ2 e† is similar.
In what follows we assume that
λ ∈ D+
inv , (4.38)
104 Chapter 4. Lines and curves in BC
|Z · W |k = λ · μ ∈ D+
inv ,
hence
Z · W ∈ Sλμ .
These properties hint that we should consider all the spheres “together” and we
set: #
SD := Sλ
λ∈D+
inv
" D := {Sλ }
S λ∈D+ inv
Sλ ◦ Sμ := Sλμ .
The unitary sphere S1 will serve as the identity in S " D (i.e., the multiplicative
neutral element), and the sphere Sλ−1 is the inverse of the sphere Sλ .
The subset
" R := {Sλ }
S λ∈R+
ϕ : λ ∈ D+ "
inv −→ Sλ ∈ SD
"R.
ϕ|R+ : λ ∈ R+ −→ Sλ ∈ S
The sets SD and S" D have proved to be multiplicative groups since the zero-divisors
were forbidden to be radii of the spheres. Let us return to these exceptions and
let us consider the sets
# #
Se,D := Sλ ; Se† ,D := Sλ .
λ∈D+
e λ∈D+†
e
+
Note that the sets D+e and De† have become multiplicative groups when a specific
multiplication was introduced in each of them. This allows us, again, to make Se,D
and Se† ,D multiplicative groups following the pattern of Section 2.7.1. If Z ∈ Sλ
and W ∈ Sμ , i.e., λ = λ1 e and μ = μ1 e, Z = β1 e with |β1 | = λ1 > 0 and W = γ1 e
with |γ1 | = μ1 > 0; then Z W := β1 γ1 e ∈ Sλ1 μ1 e = Sλμ . Hence, with this new
multiplication the set Se,D becomes a multiplicative group. The same for Se† ,D .
The analogues of S " D are the sets
Sλ Sμ := Sλμ .
The notion of limit for complex functions is well known and we will not redis-
cuss it here. Note that the formal proofs of its properties depend strongly on the
properties of the modulus of a complex number;
1
|ab| = |a| · |b| , |a + b| ≤ |a| + |b| , = 1 for a = 0 . (5.1)
a |a|
and it turns out that they allow us to repeat most of the proofs almost literally,
although one needs to take into account the fact that |Z −1 | is not always equal to
1
.
|Z|
lim Zn = Z0 ,
n→∞
which gives directly that a sequence {Zn }n∈N converges to Z0 if and only if the
corresponding coordinate sequences converge to the respective components of Z0 .
For instance, writing Zn = z1n + jz2n = a1n e + a2n e† we conclude that {Zn }n∈N
converges to Z0 = z10 + jz20 = a10 e + a2o e† if and only if {z1n }n∈N converges to
z10 and {z2n }n∈N converges to z20 , and if and only if {a1n }n∈N converges to a10
and {a2n }n∈N converges to a20 . The same happens for the other representations.
Consider now an arbitrary bicomplex convergent sequence Zn → Z0 ; if Zn =
a1n e + a2n e† , Z0 = a10 e + a20 e† and assuming that Z0 is not a zero-divisor, then
a10 = 0, a20 = 0; since a1n → a10 and a2n → a20 and these are complex sequences,
then there exists n0 ∈ N such that a1n = 0 and a2n = 0 for all n ≥ n0 . In other
words starting from n = n0 our sequence contains only invertible terms.
For the sum, product and quotient of two bicomplex sequences the usual
statements hold, and we comment on the cases of the product and of the quotient.
Let limn→∞ Zn = Z0 and limn→∞ Wn = W0 . For any n ∈ N there holds:
where we have used properties (5.2). Take M > 0 such that |Zn | < M and
|Wn | < M for all n ∈ N. Now given an arbitrary > 0 there exists N ∈ N
such that simultaneously
|Zn − Z0 | < √ and |Wn − Z0 | < √ , ∀n ≥ N .
2 2M 2 2M
Geometrically, such a ball can be seen as a bidisk in C2 , thus all such balls form
another basis in the topology τeuc .
In Section 5.1 we introduced two formally different types of convergent se-
quences. Now we realize they are the same convergence with respect to the Eu-
clidean topology but one of the definitions deals with the Euclidean basis of it
and the other deals with the bicomplex balls with non-zero-divisor radius. Thus,
in accordance with the problem we are faced, we will use one or another basis of
the topology, depending on which one of them is more appropriate.
(II) if the limit lim F (Z) exists, then the function F is bounded in a Euclidean
Z→Z0
ball with center in Z0 and it is D-bounded in a bicomplex ball with a non-
zero-divisor radius;
(III) if lim F (Z) = A ∈
/ S0 , then there exists a ball B with center in Z0 such
Z→Z0
that for all Z ∈ B, F (Z) ∈
/ S0 ;
(IV) if lim F (Z) = A, lim G(Z) = B, then the sum, the product and the
Z→Z0 Z→Z0
quotient (if B ∈
/ S0 ) have limits at Z0 and the usual formulas hold.
A bicomplex function is continuous at a point Z0 ∈ Ω ⊂ BC, if lim F (Z)
Z→Z0
exists and
lim F (Z) = F (Z0 ) .
Z→Z0
Elementary Bicomplex
Functions
Historically, a small collection of real functions was assigned the name of ele-
mentary functions: polynomials, rational functions, the exponential, trigonometric
functions; together with their inverses: the n-th root, logarithm, inverse trigono-
metric functions. Later, for their complex extensions the same name has been
preserved, although many other functions have emerged which could rightly be
called elementary as well.
The aim of this chapter is to show that the structure and the properties of
bicomplex numbers allow us to further extend to BC all those elementary functions
in such a way that the extensions keep having amazingly many properties and
features of their real and complex antecedents.
where the ak are complex numbers and where z is a complex variable. We assume
that the leading coefficient an = 0 so that the polynomial is said to have degree
n. In particular, a polynomial of degree zero is, by definition, a non-zero constant.
When all ak are real numbers, the polynomial p(z) is called a real polynomial of
one complex variable. Observe that p(z) is a real polynomial of a complex variable
if and only if
p(z) = p(z) (6.1)
for all z ∈ C. From this it follows that, for such a polynomial, if p(a) = 0, then
p(a) = 0; therefore either a is real or p has the conjugate pair of zeros a and a.
Rewriting (6.1) as p(z) = p(z), one concludes that the range p(C) of the real
polynomial p is symmetric with respect to the real axis.
One of the most remarkable properties of complex polynomials is captured by
Gauss’s Fundamental Theorem of Algebra which states that a complex polynomial
of degree n, n > 0, has exactly n zeros considering the multiplicities. The usual
proofs of this use methods of analysis or topology (it is not a result which follows
purely from the algebraic field property of the complex numbers). In particular,
we have then that if p(z) and q(z) are polynomials of degree not exceeding n and
if the equation
p(z) = q(z)
is satisfied at n + 1 distinct points, then p = q.
Notice that real polynomials of a complex variable obey the Fundamental
Theorem of Algebra, but the situation with real polynomials of a real variable is
totally different: such a polynomial of degree n > 0 might have any number of zeros
up to n. We will see what happens, in this sense, with bicomplex polynomials.
$
n
$
n
p(Z) = γk β1k e + δk β2k e† =: φ(β1 )e + ψ(β2 )e† .
k=0 k=0
S = S1 e + S2 e† ,
Their distinct roots are S1 = {i, 1 + 2i} and S2 = {0, 3i}. Then p has the following
four roots:
1 1 1 + 2i −2 + i 1 + 5i 1+i
S= i − j, 2i + j, +j , +j .
2 2 2 2 2 2
Example 6.1.2.2. Consider the polynomial
p(Z) = (1 + ji) Z 2 − (i − j) .
Example 6.1.2.3. Slightly adjusting the previous example, i.e., taking ψ(β2 ) ≡ 2,
we get the polynomial
p(Z) = (1 + ji) Z 2 + (1 − i) + j (1 − i) ,
which has no roots.
It is also important to note that a bicomplex polynomial may not have a
unique factorization into linear polynomials. For example, the polynomial p(Z) =
Z 3 − 1 has 9 zeroes. Indeed, the associated complex polynomials are
φ(β1 ) = β13 − 1, φ(β2 ) = β23 − 1.
The set of zeros of φ and ψ are, respectively:
√ √ %
1 3 1 3
S1 = β1,1 = 1, β1,2 = − + i , β1,3 = − − i
2 2 2 2
√ √ %
1 3 1 3
S2 = β2,1 = 1, β2,2 = − + i , β2,3 = − − i
2 2 2 2
and √ √
3 1 3 1 3
Z − 1 = (Z − 1) Z + − j Z + +j .
2 2 2 2
It is therefore clear from what we have indicated that bicomplex polynomials
do not satisfy the Fundamental Theorem of Algebra in its original form. At the
same time, the following results are true and summarize the comments above.
Theorem 6.1.2.4 (Analogue of the Fundamental Theorem of Algebra for bicom-
$
n
plex polynomials). Consider a bicomplex polynomial p(Z) = Ak Z k . If all the
k=0
coefficients Ak with the exception of the free term A0 = γ0 e + δ0 e† are complex
multiples of e (respectively of e† ), but δ0 = 0 (respectively γ0 = 0), then p has no
roots. In all other cases, p has at least one root.
6.1. Polynomials of a bicomplex variable 117
for all Z ∈ BC. This implies that if p(Z0 ) = 0, then p(Z 0 ) = 0, p(Z0† ) = 0,
p(Z0∗ ) = 0; now if Z0 ∈ R in this case Z0 does not have “associated” roots of p. If
Z0 is not real, then the following situations arise:
• Z0 ∈ C(i) \ R, then Z 0 = Z0 is also a root of p.
meaning that the range p(BC) of the polynomial p with real coefficients possesses
all the three symmetries generated by the three conjugations.
$n
Assume now that the coefficients of the polynomial p(Z) = Ak Z k are in
k=0
C(i), then
p(Z) = p(Z † )† (6.3)
for all Z in BC. This implies that if Z0 is a root of p, p(Z0 ) = 0, then p(Z0† ) = 0; if
Z0 is a C(i)-complex number, then Z0 does not have “associated” roots of p. But
if Z0 is not in C(i), then Z0† is also a root of p. Of course, since (6.3) is equivalent
to p(Z † ) = p(Z)† , one may conclude that the range p(BC) of such a polynomial
has the symmetry determined by the †-conjugation.
118 Chapter 6. Elementary Bicomplex Functions
As the next step, we assume that the coefficients of the polynomial p are in
C( j), then
p(Z) = p(Z) (6.4)
for all Z in BC. This implies that if Z0 is a root of p, p(Z0 ) = 0, then p(Z 0 ) = 0;
again if Z0 is a C( j)-complex number, then Z0 does not have “associated” roots
of p. But if Z0 is not in C( j), then Z 0 is also a root of p. Of course, since (6.4)
is equivalent to p(Z) = p(Z), one may conclude that the range p(BC) of such a
polynomial has the symmetry determined by the bar-conjugation.
Finally, we consider the case of hyperbolic coefficients of the polynomial p.
Here
p(Z) = p(Z ∗ )∗ (6.5)
for all Z in BC. This implies that if Z0 is a root of p, p(Z0 ) = 0, then p(Z0∗ ) = 0; if
Z0 is a hyperbolic number, then Z0 does not have “associated” roots of p. But if
Z0 is not in D, then Z0∗ = Z0 is also a root of p. Of course, since (6.5) is equivalent
to p(Z)∗ = p(Z ∗ ), one may conclude that the range p(BC) of such a polynomial
has the symmetry determined by the ∗-conjugation.
Definition 6.1.2.6. We will use the name bicomplex rational functions for functions
of the form
p(Z)
q(Z)
with two bicomplex polynomials p and q.
Such a function is well-defined for those values of Z for which the polynomial
q(Z) = Bm Z m + Bm−1 Z m−1 + · · · + B1 Z + B0 takes values in BCinv := BC \ S0 .
If both polynomials are of degree one: p(Z) = A1 Z + A0 , q(Z) = B1 Z + B0 , then
we have a fractional linear transformation. If the coefficients A1 and B1 are both
invertible, then the fractional linear transform takes the form
C2
C1 +
Z + C3
with bicomplex C1 , C2 , C3 .
the reader will be able to derive Euler’s formula eix = cos x + i sin x as a formal
identity.
The exponential function can be derived in many different ways.
If one recalls
n
1
that the Euler’s number e arises as the limit of the sequence 1 + , then one
n
x n
is tempted to consider, for any real number x, the sequence 1 + and its limit
n
which brings us to the definition of the real exponential function as
x n
exp(x) = ex := lim 1 + . (6.6)
n→∞ n
This approach has the advantage of using minimal mathematical tools; as a
matter of fact, only the properties of sequences and of their limits are necessary,
meanwhile the definitions as the sum of a convergent power series or as a solution
of a differential equation require much more elaborated techniques.
Defined by (6.6), the real exponential function preserves all the expected
properties; for example, if x ∈ N, then ex = e& · e'(· · · e). It has the property
x times
thus, it is a homomorphism of the additive group of real numbers into the multi-
plicative group R \ {0}. Since the real exponential function is monotone, then it
realizes an isomorphism of these groups.
It turns out that definition (6.6) extends to the complex numbers in the sense
that for any z ∈ C
z n
exp(z) = ez := lim 1 + . (6.8)
n→∞ n
This implies the famous Euler formula
ez = ex (cos y + i sin y)
Proof. The computation below proves that the sequence is convergent. Set as
before Z = β1 e + β2 e† . Then
n
n
n
Z β1 β2 β1 β2
1+ = 1 + e + e† = e + e † + e + e†
n n n n n
n
n
n
β1 β2 β1 β2
= 1+ e+ 1+ e† = 1+ e+ 1+ e† .
n n n n
Relying
n on the fact
that the corresponding sequences of complex numbers
n
β1 β2
1+ and 1 + are convergent to the complex exponentials eβ1 and
n n
eβ2 , respectively, we get that the limit of the right-hand side exists. What is more,
n
n
n
Z β1 β2
lim 1+ = lim 1+ e+ 1+ e† = eβ1 e + eβ2 e†
n→∞ n n→∞ n n
1 β1 i
= (e + eβ2 ) + j (eβ1 − eβ2 )
2 2
1 z1 −iz2 z1 +iz2 i (6.9)
= (e +e ) + j (ez1 −iz2 − ez1 +iz2 )
2 2
1 −iz2 i
=e z1
(e + e ) + j (e−iz2 − eiz2 )
iz2
2 2
= ez1 (cos(z2 ) + j sin(z2 )) .
which exemplifies once more the peculiarity of the different writings of bicomplex
numbers.
We pass now to the properties of this newly introduced bicomplex exponential
function.
• First we note that the bicomplex exponential is an extension to BC of the
complex exponential function: indeed, for Z = z1 + j0 ∈ C, we have that
where cosh and sinh are the classical hyperbolic cosine and sine functions,
thus arriving at the definition of the hyperbolic (in the sense of hyperbolic
numbers) exponential function.
• Note that ez1 is the complex modulus of the bicomplex number eZ and z2 is
the complex argument of the same bicomplex number eZ . The reader may
find it instructive to compare this fact with what happens in the complex
case.
• For Z = 0 = 0e + 0e† , we have: e0 = 1e + 1e† = 1.
• For any bicomplex number Z, the exponential eZ is invertible. This is because
and the exponential terms ez1 −iz2 and ez1 +iz2 are complex exponential func-
tions, so they are never zero. The inverse of eZ is
Thus, the range of the bicomplex exponential function does not contain either
the zero or any zero-divisors.
122 Chapter 6. Elementary Bicomplex Functions
• The complex formula eiπ + 1 = 0 remains valid for bicomplex numbers, but
it is complemented with its mirror image e jπ + 1 = 0.
• For any Z = β1 e + β2 e† ∈ BC and any invertible bicomplex number W =
γ1 e + γ2 e† , i.e., γ1 γ2 = 0, the equation eZ = W is equivalent to the system
eβ1 = γ1 and eβ2 = γ2 . Because γ1 γ2 = 0, it follows that there is always a
solution. Of course, this is the first step to talk about the bicomplex logarithm
which will be commented on below.
6.3. Trigonometric and hyperbolic functions of a bicomplex variable 123
• Recalling that the complex exponential function and the complex trigono-
metric functions are periodic, we obtain that
eZ = ez1 (cos(z2 ) + j sin(z2 ))
= ez1 +2πim (cos(z2 + 2πn) + j sin(z2 + 2πn))
= eZ+2π(mi+nj) ,
for integer numbers m and n. Thus the bicomplex exponential function is
periodic with bicomplex periods 2π(mi + nj). One can prove that these are
the only periods.
ejz2 + e−jz2
cos z2 = ,
2
ejz2 − e−jz2
sin z2 = .
2j
Thus we are in a position to introduce the bicomplex sine and cosine functions
which are direct extensions of their complex antecedents.
Definition 6.3.2.1. Let Z ∈ BC. We define the bicomplex cosine and sine functions
of a bicomplex variable as follows:
ekZ + e−kZ
Note that does not give the same cos Z but it gives the hy-
2
perbolic cosine of a bicomplex number. See Section 6.3.3 below.
Given Z = z1 + jz2 = β1 e + β2 e† ∈ BC, the properties of the bicomplex
exponential bring us immediately to the idempotent representation of cos Z and
sin Z:
we obtain that
One may write Z in different forms to observe how the above formulas change.
We continue with a description of some basic properties of the bicomplex
trigonometric functions.
• Since the complex sine and cosine functions are periodic with principal period
2π, then taking Z = β1 e + β2 e† and setting Zk, = (β1 + 2kπ)e + (β2 + 2π)e†
for arbitrary integers k, we have:
Thus the real number (2π)e + (2π)e† = 2π remains the principal period of
both bicomplex sine and cosine functions.
cos(β1 ) = 0, cos(β2 ) = 0.
π π
The solutions are β1 = + kπ, and β2 = + π, for k, ∈ Z. Note that β1
2 2
and β2 are never zero, so the bicomplex solutions Z to cos Z = 0 are always
hyperbolic invertible numbers. In the {1, j} basis, we get the general solution
to cos Z = 0 as
π
Z = z1 + jz2 = ((1 + k + ) + j i(k − )) , (6.16)
2
a set of hyperbolic numbers.
sin(β1 ) = 0, sin(β2 ) = 0.
The solutions are β1 = kπ, and β2 = π, for k, l ∈ Z. Note that there are
non-invertible solutions for sin Z = 0, e.g., for β1 = 0, i.e., k = 0, and β2 = 0.
In the {1, j} basis, we get the general solution for sin Z = 0 as
π
Z = z1 + jz2 = (k + + j i(k − )) ,
2
again a set of hyperbolic numbers.
• Formulas (6.15) guarantee that the usual trigonometric identities are true,
e.g., the sums and differences of angle formulas, the double angle identities,
etc. For example:
sin2 Z + cos2 Z = (sin2 (β1 ) + cos2 (β1 ))e + (sin2 (β2 ) + cos2 (β2 ))e† = 1.
126 Chapter 6. Elementary Bicomplex Functions
• Taking the argument equal to the idempotents e and e† gives again funny
formulas. Indeed, if Z = e, i.e., β1 = 1 and β2 = 0, then
† cos(1) + 1 cos(1) − 1
cos e = cos(1)e + e = −j i ,
2 2
sin(1) sin(1)
sin e = sin(1)e = −j i .
2 2
ejZ − e−jZ
tan Z = −j = tan(β1 )e + tan(β2 )e† .
ejZ + e−jZ
In a similar fashion we have
Definition 6.3.2.3. Let Z be in BC. We define the bicomplex cotangent function
of a bicomplex variable:
cos Z
cot Z := (6.18)
sin Z
whenever sin Z is invertible, i.e., both complex numbers β1 and β2 are not integer
multiples of π.
A direct computation yields:
ejZ + e−jZ
cot Z = j = cot(β1 )e + cot(β2 )e† .
ejZ − e−jZ
6.3. Trigonometric and hyperbolic functions of a bicomplex variable 127
As in the previous section, these formulas would yield the usual properties analo-
gous to hyperbolic complex functions. For example,
cosh2 Z − sinh2 Z = cosh2 (β1 ) − sinh2 (β1 ) e + cosh2 (β2 ) − sinh2 (β2 ) e† = 1 .
where ln |w| is the real logarithm of the positive number |w|, m is any integer.
Hence, for any non-zero complex number there are infinitely many complex num-
bers which can be equally called its logarithm.
We will use the following notations and names:
Ln (w) := lnm w m ∈ Z
is called the complex logarithm of w. So, Ln w is a set and, thus, it does not
generate a univalued function but rather a multivalued function, which explains
the name for lnm w: for each m it determines a function, which can be called a
logarithmic function.
Note an important property of logarithm:
where the plus in the right-hand side among the two sets means that the resulting
set is obtained by adding all the elements of the first item and all the elements of
the second item. Observe that the various branches of the logarithmic function do
not possess this property.
it follows that
z1 ∈ Ln |W |i ,
130 Chapter 6. Elementary Bicomplex Functions
is called the principal value of the (bicomplex) logarithm of W ; then the number
with two arbitrary integers m and n, is called the (m, n)-th branch of the bicomplex
logarithm; and finally the set
Ln(W ) := lnm,n (W ) m, n ∈ Z
elnm,n (Z) = elnm |Z|i +j argi (Z)+2nπj = elnm |Z|i e j argi (Z)
= |Z|i (cos(argi (Z)) + j sin(argi (Z))) = Z.
6.6. On bicomplex inverse trigonometric functions 131
for all m, n ∈ Z.
• For Z1 and Z2 two invertible bicomplex numbers, the following formula holds:
It is clear how to introduce the principal value, the (m, n)-th branch, etc., as
well as how to deal with other inverse functions. We leave this to the interested
reader, who is recommended also to consider what happens if we solve the equation
eiZ + e−iZ
cos(Z) = = W.
2
where θ is a C(i)-argument of Z, i.e., any element in ArgC(i) (Z), see (3.7); and
where ψ is a C( j)-argument of Z, i.e., any element in ArgC( j) (Z), see (3.13). But
hence
Z = |Z|i · e jθ = |Z|i · e j argC(i),m (Z) (6.24)
and
Z = |Z|j · eiψ = |Z|j · ei argC( j),n (Z) (6.25)
for any integers m and n. We will refer, sometimes, to (6.24) and (6.25) as the
(complex) exponential representations of Z. Note also that the properties of the
exponential function imply that
Z † = |Z|i · e−jθ
and
Z = |Z|j · e−iψ .
The trigonometric representation of a bicomplex number in hyperbolic terms
is allowed both for invertible numbers and for zero-divisors. Beginning with an
invertible Z we should recall that
Z = |Z|k (cos ΨZ + i sin ΨZ ) = |Z|k eiν1 · e + eiν2 · e†
Due to the various ways in which bicomplex numbers can be written, the function
F inherits analogous representations, specifically:
This leads us to
Definition 7.1.1. The complex partial derivatives of the bicomplex function F are
defined as the following limits (if they exist):
F (Z0 + h1 ) − F (Z0 )
Fz1 (Z0 ) := lim , (7.3)
h1 →0 h1
F (Z0 + jh2 ) − F (Z0 )
Fz2 (Z0 ) := lim , (7.4)
h2 →0 h2
F (Z0 + κ1 ) − F (Z0 )
Fζ1 (Z0 ) := lim , (7.5)
κ1 →0 κ1
F (Z0 + iκ2 ) − F (Z0 )
Fζ2 (Z0 ) := lim . (7.6)
κ2 →0 κ2
Similarly we introduce the hyperbolic partial derivatives: if we recall for-
mula (1.5) and write the increment as
F (Z0 + h1 ) − F (Z0 )
Fz1 (Z0 ) := lim ,
h1 ∈S
/ 0 (D), h1 →0 h1
F (Z0 + i h2 ) − F (Z0 )
Fz2 (Z0 ) := lim , (7.7)
h2 ∈S
/ 0 (D), h2 →0 h2
where with the symbol S0 (D) := S0 ∩ D we indicate the set of the hyperbolic
zero-divisors together with 0 ∈ D.
Example 7.1.2. From the point of view of the classic theory of functions of two
complex variables, one may think, because of the “symmetry” of the imaginary
7.2. The bicomplex derivative and the bicomplex derivability 137
units i and j, and thus the analogy between the corresponding complex variables,
that there is no reason to define all the complex partial derivatives (7.3)–(7.6). The
following simple example illustrates that inside the bicomplex realm the differences
are relevant. Consider to this purpose the function F : BC → BC, F (Z) = Z † . It
is immediate to see that
and that neither Fζ1 (Z0 ) nor Fζ2 (Z0 ) exist for any Z0 ∈ BC.
Remark 7.1.3. Formulas (1.6)–(1.8) suggest the introduction of six more partial
derivatives: four complex and two hyperbolic. But since there exists a direct relation
between the corresponding variables, namely,
it is easy to see that we do not get anything essentially new. The situation with
partial derivatives with respect to variables arising from the idempotent represen-
tations of bicomplex numbers is on the other hand much more interesting, and will
be discussed in detail later.
and
Remark 7.2.3. It is necessary to make a comment here. Traditionally, see e.g. [103],
p. 138 and p. 432, if h is either a real or a complex increment, the symbol o(h)
138 Chapter 7. Bicomplex Derivability and Differentiability
is used to indicate any expression of the form α(h)|h| with lim α(h) = 0. Since,
h→0
|h|
both in the real and in the complex case, the expression remains bounded when
h
h → 0, it is clear that one could replace α(h)|h| by α(h)h in the expression of o.
However, the situation is quite different in the bicomplex case. Here, the expression
|H|
is not bounded when H → 0, and therefore we need to carefully distinguish the
H
two expressions. In accordance with the usual notation, we will always use o(H)
to denote a function of the form α(H)|H|, and therefore the expression in the
previous corollary αF,Z0 (H)H is not, in general, o(H). This distinction is at the
basis of the notions of weak and strong Stoltz conditions for bicomplex functions,
which are used by G.B.Price in [56].
Remark 7.2.4. A bicomplex function F derivable at Z0 enjoys the following pro-
perty:
lim (F (Z0 + H) − F (Z0 )) = 0 . (7.10)
H ∈S
/ 0 , H→0
(C · F ) (Z0 ) = C · F (Z0 ).
F
4. If G is continuous at Z0 and G(Z0 ) ∈ S0 , then the quotient is derivable
G
at Z0 and
F F (Z0 ) · G(Z0 ) − F (Z0 ) · G (Z0 )
(Z0 ) = .
G (G(Z0 ))2
7.2. The bicomplex derivative and the bicomplex derivability 139
Proof. From Corollary 7.2.2, there exist two bicomplex functions αF,Z0 and αG,Z0
such that
lim αF,Z0 (H) = lim αG,Z0 (H) = 0
S0 H→0 S0 H→0
and
for all H ∈ S0 . The first statement about the sum or difference F ± G is ob-
tained easily by adding (or subtracting) formulas (7.11) and (7.12), and noting
that the functions αF ±G,Z0 := αF,Z0 ± αG,Z0 respect the properties from Corol-
lary 7.2.2 (which is an “if and only if” statement). Similarly, if we multiply (7.11)
by the number C, we obtain statement (2). Note that C needs not necessarily be
an invertible bicomplex number. For example, if C = c e, where c ∈ C(i), then
from (7.11) (multiplied by C) we obtain that
Now, writing
we obtain:
lim αF G,Z0 (H) = 0
S0 H→0
and
which proves that the product F · G is derivable at Z0 and the usual product
formula holds.
140 Chapter 7. Bicomplex Derivability and Differentiability
where α F ,Z0 (H) is the bicomplex function containing all the remaining terms and
G
which enjoys the property
lim α F ,Z0 (H) = 0 .
S0 H→0 G
Using again Corollary 7.2.2, the proof of the derivability of the quotient and of
the corresponding formula is finished.
Theorem 7.2.6. Let F and G be two bicomplex functions, F is defined on an open
set Ω and G is defined on F (Ω) ⊂ BC. Assume that there is a point Z0 ∈ Ω such
that W0 = F (Z0 ) is an interior point of F (Ω) and such that
F (Z) − F (Z0 ) ∈ S0 , implies Z − Z0 ∈ S0 , Z ∈ Ω.
Assume now that F is derivable at Z0 and G is derivable at W0 . Then the com-
position G ◦ F is derivable at Z0 and
(G ◦ F ) (Z0 ) = G (F (Z0 )) · F (Z0 ).
7.2. The bicomplex derivative and the bicomplex derivability 141
and
F (Z0 + H) − F (Z0 ) = F (Z0 ) · H + αF,Z0 (H)H , (7.14)
G(W0 + K) − G(W0 ) = G (W0 ) · K + αG,W0 (K)K , (7.15)
for all H, K ∈ S0 .
Denote by VW0 a neighborhood of W0 contained in F (Ω) which exists by
∗
hypothesis. Let then VW 0
be the set obtained from VW0 by eliminating the points
W for which W − W0 are zero-divisors. Let UZ∗ 0 be the set of all Z ∈ UZ0 :=
F −1 (VW0 ) such that Z − Z0 is not a zero-divisor.
Now we note that formula (7.15) is true for those invertible K such that
∗
WK := W0 + K ∈ VW0 ; such K exist by taking WK ∈ VW 0
. Then by the properties
∗
of F , for any such K there exists a ZK ∈ UZ0 such that WK = W0 + K = F (ZK ).
Now we write:
ZK = Z0 + H := Z0 + (ZK − Z0 ) ,
where H is invertible by the choice of ZK . Therefore we replace W0 +K = F (ZK ) =
F (Z0 + H), W0 = F (Z0 ) and K = F (Z0 + H) − F (Z0 ) in formula (7.15), and we
obtain:
G(F (Z0 + H)) − G(F (Z0 )) = G (F (Z0 )) · (F (Z0 + H) − F (Z0 ))
+α"G,F (Z0 ) (H) · (F (Z0 + H) − F (Z0 )) ,
where α "G,F (Z0 ) (H) := αG,F (Z0 ) (F (Z0 + H) − F (Z0 )) is a function having the
property from Corollary 7.2.2, as when K → 0 so does H → 0, in both cases along
invertible bicomplex numbers. Now we use formula (7.14) and we get:
G(F (Z0 + H)) − G(F (Z0 )) = G (F (Z0 )) · (F (Z0 ) · H + αF,Z0 (H)H)
+α"G,F (Z0 ) (H) · (F (Z0 ) · H + αF,Z0 (H)H)
=: G (F (Z0 )) · F (Z0 ) · H + αG◦F (H)H ,
where the function αG◦F (H) has obviously the desired property from Corollary 7.2.2.
The last statement of the theorem can be proved quite analogously to the
cases of real and complex functions. If one assumes beforehand the existence of the
derivative of the inverse function at Wo , then the corresponding formula becomes
a particular case of the chain rule above:
1 = (F −1 ◦ F ) (Z0 ) = (F −1 ) (F (Z0 )) · F (Z0 ) ,
142 Chapter 7. Bicomplex Derivability and Differentiability
where we used that the derivative of the identity function is one which is obvious.
Theorem 7.2.7 (The derivability of bicomplex elementary functions). All bicom-
plex elementary functions introduced in Chapter 6 are derivable at any point
Z = Z0 where they are defined. In more detail:
1. Any constant function is derivable with the derivative zero.
2. Bicomplex polynomials are derivable for any Z ∈ BC. In particular, we have:
(Z n ) = nZ n−1 .
holds for any A, B ∈ BC, due to the algebraic properties of addition and multi-
plication of bicomplex numbers. Applying this for A = Z + H and B = Z, we
obtain:
(Z + H)n − Z n
lim
S0 H→0 H
H (Z + H)n−1 + (Z + H)n−2 Z + · · · + Z n−1
= lim
S0 H→0 H
= lim (Z + H) n−1
+ (Z + H)n−2 Z + · · · + Z n−1 = nZ n−1 .
S0 H→0
7.2. The bicomplex derivative and the bicomplex derivability 143
$
n
It follows that any bicomplex polynomial p(Z) = Ak Z k is derivable at any
k=0
Z ∈ BC, and its derivative is a polynomial of degree one less, given by
$
n
p (Z) = kAk Z k−1 .
k=1
eZ+H − eZ eH − 1
= eZ · .
H H
We show now that
eH − 1
lim = 1.
S0 H→0 H
For this, let us write H = he + ke† in the idempotent representation, where
h, k ∈ C(i) are such that hk = 0. Then eH = eh e + ek e† , thus
eH − 1 e h e + e k e† eh − 1 ek − 1 †
= = e + e .
H he + ke† h k
Therefore
eH − 1 eh − 1 ek − 1
lim = lim e + lim e† = e + e† = 1 .
S0 H→0 H h→0 h k→0 k
It follows that
eZ+H − eZ eH − 1
lim = eZ · lim = eZ ,
S0 H→0 H S0 H→0 H
Since the exponential function is derivable for all Z ∈ BC, it follows that cos Z is
derivable for all Z and
Similarly the function sin Z is derivable for all Z and (sin Z) = cos Z.
144 Chapter 7. Bicomplex Derivability and Differentiability
sin Z
The bicomplex tangent function is defined as the quotient , whenever
cos Z
cos Z is invertible. Using the quotient rule, we obtain:
cos2 Z + sin2 Z 1
(tan Z) = = .
cos2 Z cos2 Z
A similar argument applies for the function cot Z.
The hyperbolic functions of a bicomplex variable are defined as
eZ + e−Z eZ − e−Z
cosh Z = , sin Z = .
2 2
Again, using that eZ is derivable for any Z, it follows that cosh Z and sinh Z are
derivable for any Z and
eZ − e−Z eZ + e−Z
(cosh Z) = = sinh Z, (sinh Z) = = cosh Z .
2 2
Recall that the (m, n)-branch bicomplex logarithmic function lnm,n , defined
for all invertible bicomplex numbers, is the inverse function of eZ (at least on a
subset of its domain of definition). Then we have:
1 = elnm,n (Z) = elnm,n (Z) · (lnm,n ) (Z) = Z · (lnm,n ) (Z) .
F (Z0 + x1 ) − F (Z0 )
lim ,
x1 →0 x1
∂F
and it coincides with (Z0 ).
∂x1
The rest of the proof follows by considering the three specific forms of the
increment H along the units i, j, k, i.e., H = iy1 , with y1 = 0; H = jx2 , with
x2 = 0; H = ky2 , with y2 = 0, and noting that all are invertible bicomplex
numbers.
Let us write the bicomplex function as F = f11 + if12 + jf21 + kf22 in terms
of its real components, which are all real functions of a bicomplex variable. An
immediate consequence of the theorem above is
Corollary 7.3.2. If F is derivable at Z0 , then the the real Jacobi matrix of F at
Z0 is of the form
⎛ ⎞
a −b −c d
⎜ b a −d −c ⎟
JZ0 [F ] := ⎜
⎝ c −d a
⎟, (7.17)
−b ⎠
d c b a
where
∂f11 ∂f12 ∂f21 ∂f22
a := = = = ,
∂x1 ∂y1 ∂x2 ∂y2
∂f11 ∂f12 ∂f21 ∂f22
b := − = =− = ,
∂y1 ∂x1 ∂y2 ∂x2
∂f11 ∂f12 ∂f21 ∂f22
c := − =− = = ,
∂x2 ∂y2 ∂x1 ∂y1
∂f11 ∂f12 ∂f21 ∂f22
d := =− =− = , (7.18)
∂y2 ∂x2 ∂y1 ∂x1
and where all the partial derivatives are evaluated at the point Z0 .
Proof. The proof relies on a direct computation using equalities (7.16) written in
terms of fk .
Remark 7.3.3. The reader should notice that the special form of the real Jacobi
matrix above encodes several Cauchy-Riemann type conditions on (certain pairs
of ) the real functions fk , a fact which we will exploit in detail below.
146 Chapter 7. Bicomplex Derivability and Differentiability
Remark 7.3.4. Every 4 × 4 matrix with real entries determines a linear transfor-
mation on R4 . Of course, not all of them remain linear when R4 is seen as BC,
i.e., not all of them are BC-linear. Those matrices which are BC-linear are of the
form (7.17). Taking into account that the entries of (7.17) are the values of the
partial derivatives at Z0 , one may conclude that (7.17) determines a BC-linear
operator acting on the tangential BC-linear module at the point Z0 . Moreover, in
the matrix (7.17) there are hidden the two linearities with respect to C(i) and C(j),
in the following sense: fixing the identifications
and
(x1 + jx2 , y1 + jy2 ) = (ζ1 , ζ2 ) ←→ (x1 , y1 , x2 , y2 )
we have two different identifications C2 (i) ↔ R4 and C2 (j) ↔ R4 . It follows that
a 4 × 4 matrix with real entries determines not only a real transformation but a
C(i)-linear one, if and only if it is of the form
⎛ ⎞
l −m u −v
⎜ m l v u ⎟
⎜ ⎟
⎝ t −s g −h ⎠ .
s t h g
Thus the structure of the matrix (7.17) includes both complex structures.
Remark 7.3.5. An easy computation shows that the real Jacobian, i.e., the deter-
minant of the matrix (7.17), is given by
det(JZ0 ) = (b + c)2 + (a − d)2 · (b − c)2 + (a + d)2 ∈ R+ .
Moreover, a direct computation yields that the expression above is nothing but
|F (Z0 )|2i 2 ,
i.e., the square of the usual modulus of the square of the complex modulus of
F (Z0 ).
A simple computation shows that det(JZ0 ) = 0 if and only if F (Z0 ) ∈ S0 ,
which is equivalent to
Note that these relations are between the real partial derivatives of the real com-
ponents of F at Z0 . Recall that the derivative of F at Z0 is given by
F (Z0 ) = a + b i + c j + d k .
Proof. As in the proof of Theorem 7.3.1, the limit of the difference quotient must
be the same regardless of the path on which H approaches zero, as long as H is
invertible. Let us choose H = h1 = h1 + j 0 → 0, and note that if Z0 = z01 + jz02 ,
then Z0 + H = (z01 + h1 ) + jz02 . Then, since F (Z0 ) exists, the following limit also
exists:
F (Z0 + h1 ) − F (Z0 )
lim = Fz1 (Z0 ).
h1 →0 h1
Similarly, taking H = j h2 → 0 we get:
F (Z0 + jh2 ) − F (Z0 )
F (Z0 ) = lim = −jFz2 (Z0 ) .
h2 →0 jh2
148 Chapter 7. Bicomplex Derivability and Differentiability
∂F ∂F
(Z0 ) = (Z0 ) = 0 , (7.23)
∂z 1 ∂z 2
i.e.,
∂f1 ∂f1 ∂f2 ∂f2
(Z0 ) = (Z0 ) = (Z0 ) = (Z0 ) = 0 , (7.24)
∂z 1 ∂z 2 ∂z 1 ∂z 2
∂ ∂
where the symbols and are the commonly used formal operations on C(i)-
∂z 1 ∂z 2
valued functions of z1 and z2 , having partial derivatives at Z0 , namely,
∂f 1 ∂f ∂f
:= +i
∂z 1 2 ∂x1 ∂y1
and
∂f 1 ∂f ∂f
:= +i .
∂z 2 2 ∂x2 ∂y2
7.3. Partial derivatives of bicomplex derivable functions 149
Proof. Let us employ equalities (7.16) involving the real partial derivatives of
F = f1 + jf2 at Z0 . The second equality in (7.16) is
∂F ∂F
(Z0 ) = −i (Z0 ) , (7.25)
∂x1 ∂y1
which is equivalent to (for simplicity we eliminate the explicit reference to Z0 )
∂f11 ∂f12 ∂f21 ∂f22 ∂f11 ∂f12 ∂f21 ∂f22
+i +j + ij = −i +i +j + ij .
∂x1 ∂x1 ∂x1 ∂x1 ∂y1 ∂y1 ∂y1 ∂y1
Because the functions fk are real, this is equivalent to the system
∂f11 ∂f12 ∂f11 ∂f12
= , =− ,
∂x1 ∂y1 ∂y1 ∂x1
∂f21 ∂f22 ∂f21 ∂f22
= , =− . (7.26)
∂x1 ∂y1 ∂y1 ∂x1
These are the real Cauchy-Riemann conditions for the complex functions f1 =
f11 + if12 and f2 = f21 + if22 , with respect to the complex variable z1 = x1 + iy1 .
In complex notation, if we write
equality (7.25) in terms of the complex differential
1 ∂ ∂ ∂
operator = +i , it becomes equivalent to the first part of (7.23).
2 ∂z 1 ∂x1 ∂y1
We repeat this reasoning for the equality
∂F ∂F
−j (Z0 ) = ij (Z0 ), (7.27)
∂x2 ∂y2
which, after division by −j, is equivalent to the second equality in (7.23). This
leads to the conclusion that f1 and f2 verify the real Cauchy-Riemann system
with respect to the variable z2 at Z0 .
Of course, this corollary shows the relation between bicomplex derivability
and classical holomorphy in two complex variables.
Let us now express the bicomplex variable as Z = ζ1 +i ζ2 , and the bicomplex
function as F = ρ1 + iρ2 , where ρ1 = f11 + jf21 and ρ2 = f12 + jf22 are C(j)-valued
functions. We prove the following
Theorem 7.3.8. Consider a bicomplex function F derivable at Z0 . Then
1. The C(j)-complex partial derivatives Fζ (Z0 ) exist, for = 1, 2.
2. The complex partial derivatives above verify the equality:
Proof. In the definition of F (Z0 ) we consider the limit of the difference quotient
as H → 0 first through invertible values of the form H = κ1 + i0, and then
through values of the form H = 0 + iκ2 , where κ1 , κ2 ∈ C(j). A computation as
in Theorem 7.3.6 leads to the existence of the complex partial derivatives of F at
Z0 with respect to the variables ζ1 , ζ2 ∈ C(j), and one can show that they verify
the equation:
F (Z0 ) = Fζ1 (Z0 ) = −iFζ2 (Z0 ) , (7.30)
where
Fζ1 (Z0 ) = ρ1,ζ1 (Z0 ) + iρ2,ζ1 (Z0 ), Fζ2 (Z0 ) = ρ1,ζ2 (Z0 ) + iρ2,ζ2 (Z0 ) .
Now it is immediate to see that the equality (7.30) is equivalent to the sys-
tem (7.29).
Corollary 7.3.9. If F = ρ1 + iρ2 is bicomplex derivable at Z0 , then the real com-
ponents of the functions ρ1 = f11 + jf21 and ρ2 = f12 + jf22 verify the usual real
Cauchy-Riemann system (at Z0 ) associated to each complex variable ζ1 = x1 + jx2
and ζ2 = y1 + jy2 ; in complex notation this is equivalent to
∂F ∂F
(Z0 ) = ∗ (Z0 ) = 0 , (7.31)
∂ζ1∗ ∂ζ2
i.e.,
∂ρ1 ∂ρ1 ∂ρ2 ∂ρ2
(Z0 ) = ∗ (Z0 ) = ∗ (Z0 ) = ∗ (Z0 ) = 0 . (7.32)
∂ζ1∗ ∂ζ2 ∂ζ1 ∂ζ2
Here
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
∗ := +j , ∗ := +j .
∂ζ1 2 ∂x1 ∂x2 ∂ζ2 2 ∂y1 ∂y2
Proof. The following equalities from (7.16) among the real partial derivatives of
F
∂F ∂F ∂F ∂F
(Z0 ) = −j (Z0 ), −i (Z0 ) = ij (Z0 ) (7.33)
∂x1 ∂x2 ∂y1 ∂y2
are equivalent to (7.31).
Remark 7.3.10. We showed in Example 7.1.2 that it is possible for a bicomplex
function to have partial derivatives with respect to z1 , z2 , and not to have partial
derivatives with respect to ζ1 , ζ2 . Now we see that if the function is bicomplex
derivable, then such a situation is not possible: a bicomplex derivable function has
complex partial derivatives with respect to the C(i)-complex variables z1 and z2
as well as with respect to the C( j)-variables ζ1 and ζ2 . Moreover, formulas (7.20)
and (7.28) show that such derivatives are related by
∂F ∂F ∂F ∂F
(Z0 ) = k (Z0 ), (Z0 ) = −k (Z0 ) , (7.38)
∂x1 ∂y2 ∂y1 ∂x2
Remark 7.3.13. Appealing again to formulas (1.6)–(1.8) which deal with the com-
plex variables w1 , w2 , ω1 , ω2 and the hyperbolic variables w1 , w2 , one may wonder:
what about the Cauchy-Riemann conditions with respect to the corresponding par-
tial derivatives? Remark 7.1.3 explains how they can be obtained directly from the
previous statements. We omit the details.
Thus for a BC-holomorphic function F all the conclusions made in this section
hold in the whole domain. Theorem 7.3.6 says that F is holomorphic with respect
to z1 for any z2 fixed and F is holomorphic with respect to z2 for any z1 fixed.
Thus, see for instance [39, pages 4-5], F is holomorphic in the classical sense of two
complex variables. This implies immediately many quite useful properties of F , in
particular, it is of class C ∞ (Ω), and the reader may compare this with Remark 7.2.4
where we were able, working with just one point, not with a domain, to state a
weakened continuity at the point.
∂F ∂F
F (Z + H) − F (Z) = (Z)h11 + (Z)h12
∂x1 ∂y1
∂F ∂F
+ (Z)h21 + (Z)h22 + o(H) , (7.39)
∂x2 ∂y2
h1 + h1 h1 − h 1
h11 = , h12 = ,
2 2i
h2 + h2 h2 − h 2
h21 = , h22 = .
2 2i
Using this in (7.39) and grouping adequately, we obtain:
1 ∂F ∂F
F (Z + H) − F (Z) = (Z) − i (Z) · h1
2 ∂x1 ∂y1
1 ∂F ∂F 1 ∂F ∂F
+ (Z) + i (Z) · h1 + (Z) − i (Z) · h2
2 ∂x1 ∂y1 2 ∂x2 ∂y2
1 ∂F ∂F
+ (Z) + i (Z) · h2 + o(H) .
2 ∂x2 ∂y2
If we employ the C(i)-complex variables z1 = x1 + iy1 and z2 = x2 + iy2 , and the
usual complex differential operators
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
:= −i , := +i ,
∂z1 2 ∂x1 ∂y1 ∂z 1 2 ∂x1 ∂y1
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
:= −i , := +i , (7.40)
∂z2 2 ∂x2 ∂y2 ∂z 2 2 ∂x2 ∂y2
we obtain the formula
∂F ∂F
F (Z + H) − F (Z) = (Z) · h1 + (Z) · h1
∂z1 ∂z 1
∂F ∂F
+ (Z) · h2 + (Z) · h2 + o(H) . (7.41)
∂z2 ∂z 2
We emphasize that (7.41) does not express a new notion; indeed, this is simply
the condition of real differentiability for a C 1 -bicomplex function, although it is
now expressed in C(i)-complex terms.
Note that in Section 7 we introduced the symbols Fz1 (Z) and Fz2 (Z) in-
∂F ∂F
stead of the symbols (Z) and (Z) because the former are complex partial
∂z1 ∂z2
derivatives, defined, as usual, as limits of suitable difference quotients, meanwhile
the latter indicates well known operators acting on C 1 -functions. The relationship
between these two notions is clarified by the following definition and theorem.
Definition 7.4.1. A bicomplex C 1 -function F is called C(i)-complex differentiable
if
F (Z + H) − F (Z) = Fz1 (Z) · h1 + Fz2 (Z) · h2 + o(H) .
Theorem 7.4.2. A C 1 -bicomplex function F is C(i)-complex differentiable if and
only if both its components f1 , f2 are holomorphic functions in the sense of two
complex variables.
154 Chapter 7. Bicomplex Derivability and Differentiability
∂
Proof. The partial derivative Fz1 (Z) exists in Ω if and only if the operator
∂z 1
∂
(which one can think of as a dual to the operator ) annihilates the function
∂z1
F ; that is, if and only if F is holomorphic as a function of z1 ; this can be proved
by taking h2 = 0 and h1 = 0 in (7.41). What remains is
∂F ∂F
F (Z + H) − F (Z) = F (Z + h1 ) − F (Z) = (Z) · h1 + (Z) · h1 + o(H) ,
∂z1 ∂z 1
h1
and since has no limit when h1 → 0, then we conclude that Fz1 (Z) exists if
h1
∂F ∂F
and only if (Z) = 0. In this case, Fz1 (Z) = (Z).
∂z 1 ∂z1
Similarly, the partial derivative Fz2 (Z) exists in Ω if and only if the operator
∂
annihilates the function F ; this is because we can take now h1 = 0, h2 = 0
∂z 2
∂F
in (7.41). Therefore Fz2 (Z) = (Z).
∂z2
Finally, we can assume both conditions
∂F ∂F
(Z) = (Z) = 0
∂z 1 ∂z 2
∂F ∂F
F (Z + H) − F (Z) = (Z)κ1 + (Z)κ1
∂ζ1 ∂ζ 1
∂F ∂F
+ (Z)κ2 + (Z)κ2 + o(H) , (7.43)
∂ζ2 ∂ζ 2
∂
where the expressions (and similar) are the usual complex differential op-
∂ζ1
erators in C(j). Again, this simply represents the real differentiability of a C 1 -
bicomplex function in C(j)-complex terms.
7.4. Real differentiability and derivability 155
While real differentiability uniquely defines the coefficients in (7.48), one may think
that if we had begun with another writing of Z and H, then formula (7.48) would
be different, that is, other operators would have appeared in it. Direct computa-
tions however confirm that no matter in which form we write the functions and the
variables (recall that the operators in (7.47) are given in terms of C(i)-complex dif-
∂ ∂ ∂ ∂
ferential operators) the operators , †
, , and in the right-hand side
∂Z ∂Z ∂Z ∂Z ∗
of (7.48) are uniquely defined. But this requires us to clarify what is meant by this
“uniqueness”. These are the same operators but only when acting on bicomplex
functions, without taking into account any concrete intrinsic substructure in BC.
For instance, if the functions are considered C2 (i)-valued or C2 (j)-valued, then the
operators are of course different; this is just because they act on objects of differ-
ent nature. For this reason, one should be careful when working with bicomplex
functions and operators, having in mind all the time what is exactly the structure
of BC which is of interest for a specific goal.
We write here each operator in the various possible ways:
∂ 1 ∂ ∂ 1 ∂ ∂
= −j = −i
∂Z 2 ∂z1 ∂z2 2 ∂ζ1 ∂ζ2
1 ∂ ∂ 1 ∂ ∂
= −i = +k
2 ∂z1 ∂z2 2 ∂w1 ∂w2
1 ∂ ∂ 1 ∂ ∂
= +k = +j
2 ∂ω1 ∂ω2 2 ∂w1 ∂w2
1 ∂ ∂ ∂ ∂
= −i −j +k , (7.49)
4 ∂x1 ∂y1 ∂x2 ∂y2
∂ 1 ∂ ∂ 1 ∂ ∂
= + j = − i
∂Z † 2 ∂z1 ∂z2 2 ∂ζ1∗ ∂ζ2∗
1 ∂ ∂ 1 ∂ ∂
= − i = − k
2 ∂z1 ∂z2 2 ∂w1 ∂w2
1 ∂ ∂ 1 ∂ ∂
= − k = + j
2 ∂ω1∗ ∂ω2∗ 2 ∂w1 ∂w2
1 ∂ ∂ ∂ ∂
= −i +j −k , (7.50)
4 ∂x1 ∂y1 ∂x2 ∂y2
∂ 1 ∂ ∂ 1 ∂ ∂
= −j = +i
∂Z 2 ∂z 1 ∂z 2 2 ∂ζ1 ∂ζ2
1 ∂ ∂ 1 ∂ ∂
= + i = − k
2 ∂z1 ∂z2 2 ∂w1 ∂w2
1 ∂ ∂ 1 ∂ ∂
= −k = − j
2 ∂ω1 ∂ω2 2 ∂w1 ∂w2
158 Chapter 7. Bicomplex Derivability and Differentiability
1 ∂ ∂ ∂ ∂
= +i −j −k , (7.51)
4 ∂x1 ∂y1 ∂x2 ∂y2
and
∂ 1 ∂ ∂ 1 ∂ ∂
= +j = +i ∗
∂Z ∗ 2 ∂z 1 ∂z 2 2 ∂ζ1∗ ∂ζ2
1 ∂ ∂ 1 ∂ ∂
= +i = +k
2 ∂z1 ∂z2 2 ∂w1 ∂w2
1 ∂ ∂ 1 ∂ ∂
= + k = − j
2 ∂ω1∗ ∂ω2∗ 2 ∂w1 ∂w2
1 ∂ ∂ ∂ ∂
= +i +j +k . (7.52)
4 ∂x1 ∂y1 ∂x2 ∂y2
∂
Again, take a bicomplex function F and apply to it, say, the operator ; the
∂Z ∗
∂F
resulting function is a bicomplex function. But if we omit the bicomplex
∂Z ∗
structure and consider F to be a mapping from C2 (i) to C2 (i), then such an F
does understand already what the action of the operator
⎛ ⎞
∂ ∂
−
1 ∂ ∂ 1 ⎜ ∂z 1 ∂z 2 ⎟
+j = ⎜ ⎝
⎟
2 ∂z 1 ∂z 2 2 ∂ ∂ ⎠
∂z 2 ∂z 1
1 ∂ ∂
means, but it does not understand the action of the operator + i .
2 ∂ζ1∗ ∂ζ2∗
As a consequence of the previous discussion, we obtain the following result:
Theorem 7.4.3. Let F ∈ C 1 (Ω, BC): if F is BC-holomorphic, then
∂F ∂F ∂F
(Z) = (Z) = (Z) = 0 . (7.53)
∂Z † ∂Z ∂Z ∗
holds on Ω.
Proof. Since F is BC-holomorphic, formula (7.9) holds for all H ∈ / S0 . But F is
a C 1 -function, hence (7.48) holds as well for any H = 0, thus both formulas hold
for non-zero-divisors. Then (7.53) follows directly by recalling that both (7.9)
and (7.48) are unique representations for a given function F , and by comparing
them.
Remark 7.4.4. As we will show later the converse of this result is true as well, but
we need some additional steps before we can prove it.
In order to have more consistency with the previous reasonings of this section
and in analogy with the cases of functions of real or complex variables, we introduce
the following definition.
7.4. Real differentiability and derivability 159
∂F ∂F ∂F
= = = 0.
∂Z † ∂Z ∂Z ∗
For the operators involved we use the appropriate representations from the table
∂
in the previous section. Using such a representation for the operators and
∂Z
∂
leads to the system
∂Z ∗
∂F ∂F ∂F ∂F
−j = 0, +j = 0,
∂z 1 ∂z 2 ∂z 1 ∂z 2
∂F ∂F
implying that = 0 = . The latter is equivalent to the holomorphy, in
∂z 1 ∂z 2
the sense of complex functions of two C(i)-complex variables, of the components
f1 , f2 of the function F . Thus F can be seen as a holomorphic mapping from
∂F
Ω ⊂ C2 (i) → C2 (i). But we still have more information. Since = 0, then
∂Z †
F = f1 + jf2 verifies
∂F 1 ∂ ∂
(Z) = +j (f1 + jf2 )(Z)
∂Z † 2 ∂z1 ∂z2
1 ∂f1 ∂f2 ∂f2 ∂f1
= − (Z) + j (Z) + (Z) = 0.
2 ∂z1 ∂z2 ∂z1 ∂z2
160 Chapter 7. Bicomplex Derivability and Differentiability
Thus
∂f1 ∂f2 ∂f1 ∂f2
= , =− , (7.55)
∂z1 ∂z2 ∂z2 ∂z1
that is, the complex partial derivatives of the holomorphic functions f1 , f2 are not
independent; they are tied by the Cauchy-Riemann type conditions (7.55). We
reformulate the reasoning as
Proposition 7.5.1. A function F = f1 + jf2 : Ω ⊂ BC → BC is BC-holomorphic if
and only if, seen as a mapping from Ω ⊂ C2 (i) → C2 (i), it is a holomorphic map-
ping with its components related by the Cauchy-Riemann type conditions (7.55).
In other words, the theory of bicomplex holomorphic functions can be seen
as a theory of a proper subset of holomorphic mappings in two complex variables.
Each equation in Theorem 7.4.3 plays a different role: two of them together guar-
antee the holomorphy of the C(i)-complex components and the third one provides
the relation between them.
∂
It is worth noting that the operator arises in the works of J. Ryan [69] on
∂Z †
complex Clifford analysis as a Cauchy-Riemann operator which is defined directly
on holomorphic mappings with values in a complex Clifford algebra.
Next, take a bicomplex holomorphic function F in the form F = g1 + ig2 ,
where g1 and g2 take values in C(j) and we write now Z = ζ1 + iζ2 , then the
corresponding differential operators are:
∂ 1 ∂ ∂
= − i ,
∂Z † 2 ∂ζ1∗ ∂ζ2∗
∂ 1 ∂ ∂
= +i ,
∂Z 2 ∂ζ1 ∂ζ2
∂ 1 ∂ ∂
= + i .
∂Z ∗ 2 ∂ζ1∗ ∂ζ2∗
Using again Theorem 7.4.3 and Remark 7.4.4 we have that to be a bicomplex
holomorphic function means for F that
∂F 1 ∂F ∂F
= − i ∗ = 0,
∂Z † 2 ∂ζ1∗ ∂ζ2
∂F 1 ∂F ∂F
= +i = 0,
∂Z 2 ∂ζ1 ∂ζ2
∂F 1 ∂F ∂F
∗
= ∗ + i ∗ = 0.
∂Z 2 ∂ζ1 ∂ζ2
The first and the third equations together give, again, that the components g1 and
g2 of F are holomorphic functions of the C(j)-complex variables ζ1 and ζ2 , while
the second equation gives:
∂g1 ∂g2 ∂g1 ∂g2
= =− . (7.56)
∂ζ1 ∂ζ2 ∂ζ2 ∂ζ1
7.5. Bicomplex holomorphy versus holomorphy in two variables 161
∂ 1 ∂ ∂
= − i ,
∂Z † 2 ∂z1 ∂z2
∂ 1 ∂ ∂
= + i ,
∂Z 2 ∂z1 ∂z2
∂ 1 ∂ ∂
= + i .
∂Z ∗ 2 ∂z1 ∂z2
Using again Theorem 7.4.3 and Remark 7.4.4 we have that to be a BC-holomorphic
function means for F that
∂F 1 ∂F ∂F
= − i = 0,
∂Z † 2 ∂z1 ∂z2
∂F 1 ∂F ∂F
= + i = 0,
∂Z 2 ∂z1 ∂z2
∂F 1 ∂F ∂F
= + i = 0.
∂Z ∗ 2 ∂z1 ∂z2
The first and second equations together imply that the components u1 and u2 of
F are holomorphic functions of hyperbolic variables z1 and z2 , meanwhile the last
equation gives the Cauchy-Riemann type conditions:
They look exactly as their antecedent in one complex variable, but this is a totally
different thing: we deal here with D-valued functions of two hyperbolic variables
and with hyperbolic partial derivatives.
Remark 7.6.1. Note that the above calculations show that, as is well known, the
bicomplex function F of class C 1 , seen as a mapping from C2 (i) → C2 (i), is holo-
∂F
morphic with respect to βq (q = 1, 2) if and only if (Z) = 0 in Ω. Note that if
∂β q
F is a bicomplex function, and we express it in cartesian coordinates, it turns out
that F is BC-holomorphic if and only if its components are holomorphic as func-
tions of two complex variables and satisfy a Cauchy-Riemann type relation between
them. As we will show later, this is definitely not the case when we express F in
the idempotent representation. In this case, BC-holomorphy will be equivalent to
the requirement that each component is a holomorphic function of a single complex
variable and there are no relations between the components.
We emphasize that now we are considering the identification between BC
and C2 (i),
Z = β1 e + β2 e† ←→ (β1 , β2 ) ∈ C2 (i) ,
where however the basis in C2 (i) is not the canonical basis {1, j}, but rather the
idempotent basis {e, e† }.
For the next step recall the formulas
H = η 1 e + η 2 e† , H † = η 2 e + η 1 e† ,
H = η 2 e + η 1 e† , H ∗ = η 1 e + η 2 e† ,
which imply that
η1 = He + H † e† , η 1 = H ∗ e + He† ,
η2 = H † e + He† , η 2 = He + H ∗ e† .
The condition of real differentiability after substitutions becomes:
F (Z + H) − F (Z)
∂F ∂F ∂F ∂F
=H (Z)e + (Z)e† + H † (Z)e + (Z)e†
∂β1 ∂β2 ∂β2 ∂β1 (7.61)
∂F ∂F ∂F ∂F
+H (Z)e + (Z)e† + H ∗ (Z)e + (Z)e† + o(H) .
∂β 2 ∂β 1 ∂β 1 ∂β 2
Note that the expressions in the parentheses are not, yet, the idempotent forms
of anything, since the coefficients of e and e† are bicomplex numbers, not C(i)-
complex numbers. Thus we are required to make one more step. Using the formula
F = G1 e + G2 e† we arrive at
F (Z + H) − F (Z)
∂G1 ∂G2 † † ∂G1 ∂G2 †
=H (Z)e + (Z)e + H (Z)e + (Z)e
∂β1 ∂β2 ∂β2 ∂β1
∂G1 ∂G2 † ∗ ∂G1 ∂G2 †
+H (Z)e + (Z)e + H (Z)e + (Z)e + o(H) .
∂β 2 ∂β 1 ∂β 1 ∂β 2
(7.62)
164 Chapter 7. Bicomplex Derivability and Differentiability
This formula is valid for any F in C 1 (Ω), so let us analyze how BC-holomorphic
functions are singled out among those of class C 1 .
Theorem 7.6.2. The C 1 -function F is BC-holomorphic if and only if the three
bicomplex coefficients of H † , H and H ∗ in (7.62) are all zero for any Z in Ω.
Proof. The if direction follows as in Theorem 7.4.3. Specifically, since F is BC-
holomorphic, formula (7.9) holds for all H ∈ / S0 . But F is a C 1 -function, hence
(7.62) holds as well for any H = 0, thus both formulas hold for non-zero-divisors.
Then the result follows directly by recalling that both (7.9) and (7.62) are unique
representations for a given function F , and by comparing them.
In order to prove the only if, it is helpful to write explicitly the meaning of
the vanishing of these coefficients, namely:
∂G1 ∂G2
(Z)e + (Z)e† = 0 ,
∂β2 ∂β1
∂G1 ∂G2
(Z)e + (Z)e† = 0 ,
∂β 2 ∂β 1
∂G1 ∂G2
(Z)e + (Z)e† = 0 . (7.63)
∂β 1 ∂β 2
Now note that the second and the third equations, because of the independence
of e and e† , impose that G1 and G2 are C(i)-valued holomorphic functions of the
complex variables β1 , β2 and thus they have authentic complex partial derivatives.
What is more, the first equation in (7.63) says that one of the partial derivatives
∂G1 ∂G2
of each G1 and G2 is identically zero: (Z) = 0, (Z) = 0 for any Z ∈ Ω.
∂β2 ∂β1
Hence, using (7.58) and (7.59), G1 is a holomorphic function of the single variable
β1 ∈ Ω1 and G2 is a holomorphic function of the single variable β2 ∈ Ω2 . We now
want to show that these equations imply that F is BC-holomorphic. But in fact,
because of these equations, we have that for any invertible H there holds:
Remark 7.6.4. The functions G1 and G2 are independent in the sense that there
are no Cauchy-Riemann type conditions relating them.
We are in a position now to prove that the converse to Theorem 7.4.3 is true
as well.
Theorem 7.6.5. Given F ∈ C 1 (Ω, BC), then condition (7.53) implies that F is
BC-holomorphic.
Proof. If (7.53) holds, then a direct computation shows that all the three formulas
in (7.63) are true, and by Theorem 7.6.2 F is BC-holomorphic.
The direct computation mentioned above is quite useful and instructive and
we will perform it later.
Taking into account the relations between β1 , β2 and the cartesian compo-
nents z1 , z2 , we have also that
This implies that a BC-holomorphic function has derivatives of any order and
(n) (n)
F (n) (Z) = G1 (β1 ) e + G2 (β2 ) e†
(n) (n)
= G1 (Ze + Z † e† ) e + G2 (Z † e + Ze† ) e† .
Remark 7.6.7. Although formula (7.62) is quite similar to formula (7.48) its con-
sequences for the function F are paradoxically different: while formula (7.48) has
allowed us to conclude that the cartesian components f1 , f2 are holomorphic func-
tions of two complex variables which are not independent, formula (7.62) explains
to us that the idempotent components G1 , G2 are usual holomorphic functions of
one complex variable which are, besides, independent.
166 Chapter 7. Bicomplex Derivability and Differentiability
" := Ω1 · e +
But the right-hand side of the latter is well-defined on the wider set Ω
†
Ω2 · e ⊃ Ω (in general, this inclusion is proper), with the notations as in (7.58)
and (7.59). Moreover, by Theorem 7.6.3 the function F" defined by
Remark 7.6.12. Using the results of this section, we can give the proofs of the
statements of Theorem 7.2.7 in terms of the idempotent writing of the bicomplex
elementary functions. For example, to prove that the bicomplex exponential func-
tion F (Z) = eZ is BC-holomorphic, we write:
1 = x 1 + y2 , m1 = y1 − x2 , 2 = x 1 − y2 , m2 = y1 + x2 , (7.64)
or, equivalently, by
⎛ ⎞
1 0 0 1
⎜ 0 1 −1 0 ⎟
(1 , m1 , 2 , m2 )t := ⎜
⎝ 1 0
⎟ · (x1 , y1 , x2 , y2 )t . (7.65)
0 −1 ⎠
0 1 1 0
The 4 × 4 matrix from the right-hand side of (7.65) has determinant equal 4, so
it is invertible, hence
⎛ ⎞
1 0 1 0
1⎜ 0 1 0 1 ⎟
(x1 , y1 , x2 , y2 )t := ⎜ ⎟ · (1 , m1 , 2 , m2 )t , (7.66)
2 ⎝ 0 −1 0 1 ⎠
1 0 −1 0
that is,
1 + 2 m1 + m2 m2 − m1 1 − 2
x1 = , y1 = , x2 = , y2 = . (7.67)
2 2 2 2
Altogether we have an isomorphism of linear spaces φ : R4x → R4 defined by
φ
(x1 , y1 , x2 , y2 ) −→ (1 , m1 , 2 , m2 ) (7.68)
168 Chapter 7. Bicomplex Derivability and Differentiability
with the inverse linear map φ−1 : R4 → R4x given by (7.66). Because φ is a linear
isomorphism, the matrix is the Jacobi matrix of φ, so we denote it by Jx [φ]. Note
that considering the matrix transpose of (7.65) we can write:
= x · Jx [φ]t .
x = · J [φ−1 ]t .
Remark 7.7.1. Because the determinant of Jx [φ] is 4, then φ does not preserve
Euclidean distances, i.e., it is not an isometry between R4x and R4 : if the Euclidean
norm of x is 1, then the Euclidean norm of is
√ √ √
|| = 21 + m21 + 22 + m22 = 2 x21 + y12 + x22 + y22 = 2|x| = 2 .
B −→ A := Wφ ◦ B ◦ Wφ−1 ,
7.7. Cartesian versus idempotent representations in BC-holomorphy 169
and
The above general reasoning is specified now for the case of our interest,
namely, for F(V ) = C ∞ (V ) and F(U ) = C ∞ (U ) where the operators of partial
derivatives act with respect to the corresponding variables.
Let g be a function in C ∞ (V ), then g(φ(x)) is in C ∞ (U ) and the chain rule
gives:
∂(Wφ [g]) ∂
(x) = [g(φ(x))]
∂x1 ∂x1
∂g ∂1 ∂g ∂m1
= (φ(x)) · + (φ(x)) ·
∂1 ∂x1 ∂m1 ∂x1
∂g ∂2 ∂g ∂m2
+ (φ(x)) · + (φ(x)) ·
∂2 ∂x1 ∂m2 ∂x1
+ , + ,
∂g ∂g
= Wφ (x) + Wφ (x)
∂1 ∂2
+ ,
∂g ∂g
= Wφ + (x) .
∂1 ∂2
Since both the function g and the variable x are arbitrary, we obtain:
∂ ∂ ∂
◦ Wφ = Wφ ◦ + .
∂x1 ∂1 ∂2
Similar computations are made for the other variables. Summarizing, we get:
∂ ∂ ∂
= Wφ ◦ + ◦ Wφ−1 ,
∂x1 ∂1 ∂2
∂ ∂ ∂
= Wφ ◦ + ◦ Wφ−1 ,
∂y1 ∂m1 ∂m2
∂ ∂ ∂
= Wφ ◦ − ◦ Wφ−1 ,
∂x2 ∂m2 ∂m1
∂ ∂ ∂
= Wφ ◦ − ◦ Wφ−1 . (7.71)
∂y2 ∂1 ∂2
170 Chapter 7. Bicomplex Derivability and Differentiability
Briefly, we have established what the partial derivatives with respect to the canon-
ical cartesian coordinates turn out to be when one makes the change of variables
passing to the idempotent coordinates.
Considering the usual gradient operators in R4x and R4 ,
∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
∇x = , , , and ∇ = , , ,
∂x1 ∂y1 ∂x2 ∂y2 ∂1 ∂m1 ∂2 ∂m2
If F is a bicomplex function, seen as a mapping F = (f11 , f12 , f21 , f22 ) from R4x to
R4x , then we get a relation between the Jacobi matrices of F in the two coordinates
systems considered: - .
Jx [F ] = J Wφ−1 [F ] · Jx [φ] . (7.73)
It is useful to have explicitly the reciprocal relations. Let us note that from (7.71)
we obtain:
∂ ∂ ∂
Wφ−1 ◦ ◦ Wφ = + ,
∂x1 ∂1 ∂2
∂ ∂ ∂
Wφ−1 ◦ ◦ Wφ = − ,
∂y2 ∂1 ∂2
which leads to
∂ 1 ∂ ∂
= Wφ−1 ◦ + ◦ Wφ , (7.74)
∂1 2 ∂x1 ∂y2
∂ 1 ∂ ∂
= Wφ−1 ◦ − ◦ Wφ . (7.75)
∂2 2 ∂x1 ∂y2
Similarly, we have:
∂ 1 ∂ ∂
= Wφ−1 ◦ − ◦ Wφ ,
∂m1 2 ∂y1 ∂x2
∂ 1 ∂ ∂
= Wφ−1 ◦ + ◦ Wφ . (7.76)
∂m2 2 ∂y1 ∂x2
Now for a bicomplex function G, seen again as a mapping from R4 to R4 , we get:
The reader should immediately notice that the formula above is consistent with (7.73)
for F = Wφ [G] and using the fact that Jx [φ] and J [φ−1 ] are matrices inverse to
each other.
In conclusion: if we want to translate a differential expression written in the
cartesian coordinates (x1 , y1 , x2 , y2 ) into an expression with the idempotent coor-
dinates (1 , m1 , 2 , m2 ), then we use formulas (7.71); if we start with an expression
given in the idempotent coordinates, then (7.74), (7.75) and (7.76) give its equiva-
lent in the cartesian coordinates. In other words, we have obtained direct relations
between the differential operators acting on two copies of R4 , one copy with the
cartesian coordinates and another with the idempotent coordinates.
The next step is to extend the ideas above onto the C(i)-complex differential
operators (7.40) in the complex variables z1 and z2 . Writing now the C(i)-complex
idempotent coordinates β1 := 1 + im1 and β2 := 2 + im2 , the usual associated
C(i)-complex differential operators are:
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
:= −i , := −i ,
∂β1 2 ∂1 ∂m1 ∂β2 2 ∂2 ∂m2
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
:= +i , := +i . (7.79)
∂β 1 2 ∂1 ∂m1 ∂β 2 2 ∂2 ∂m2
where
⎛ ⎞ ⎛ ⎞
1 i 0 0 1 1 0 0
⎜ 1 −i 0 0 ⎟ 1⎜ −i i 0 0 ⎟
M := ⎜
⎝ 0 0
⎟, M −1 := ⎜ ⎟.
1 i ⎠ 2⎝ 0 0 1 1 ⎠
0 0 1 −i 0 0 −i i
Now, if we want to derive the formulas relating, say, the differential operators
in z with the ones in , we just have to combine the relations (7.80) above with
the formulas (7.72) and (7.77):
∇z = ∇x · M −1 = Wφ ◦ ∇ · Jx [φ] ◦ Wφ−1 · M −1
= Wφ ◦ ∇ · Jx [φ] · M −1 ◦ Wφ−1 . (7.81)
172 Chapter 7. Bicomplex Derivability and Differentiability
we incor-
If we want now to express the operators in z in terms of the ones in β,
porate in (7.81) the relation between the gradient in and the one in β:
∇z = Wφ ◦ ∇ · Jx [φ] · M −1 ◦ Wφ−1
= Wφ ◦ ∇β · M · Jx [φ] · M −1 ◦ Wφ−1 .
γ1 = 1 + j(−m1 ), γ2 = 2 + jm2 .
Notice that the computations will not yield the same formulas, since one has to
start with a different linear isomorphism φ, " for which the defining matrix is ob-
tained from (7.65) by multiplying the second column by (−1). For example, the
formulas for the bicomplex differential operators in terms of the C(j)-complex
idempotent operators are:
∂ ∂ ∂ †
= Wφ ◦ e+ e ◦ Wφ−1 ,
∂Z ∂γ1 ∂γ2
∂ ∂ ∂ †
= W φ ◦ e + e ◦ Wφ−1 ,
∂Z † ∂γ1∗ ∂γ2∗
∂ ∂ ∂ †
= Wφ ◦ e+ e ◦ Wφ−1 ,
∂Z ∂γ2 ∂γ1
∂ ∂ ∂ †
= W φ ◦ e + e ◦ Wφ−1 .
∂Z ∗ ∂γ2∗ ∂γ1∗
174 Chapter 7. Bicomplex Derivability and Differentiability
we obtain: ⎛ ⎞
e e† e† e
1⎜ −ie −ie† ie† ie ⎟
Jx [φ] · T = ⎜ ⎟. (7.88)
2 ⎝ e† e e e† ⎠
−ie† −ie ie ie†
∂
For example, if we want to write in the real idempotent coordinates, we get
∂Z
from above:
∂ 1 ∂ ∂ ∂ † ∂ †
= Wφ ◦ e−i e+ e −i e ◦ Wφ−1 .
∂Z 2 ∂1 ∂m1 ∂2 ∂m2
where F̂ = Wφ−1 [F ].
Remark 7.7.3. Let us analyze in more detail the crucial difference between the
complex differential operators in the z and β variables. For example, if we look
side-by-side at the formulas:
∂ ∂ ∂
= + ,
∂z1 ∂Z ∂Z †
∂ −1 ∂ ∂ †
= Wφ ◦ ·e+ · e ◦ Wφ ,
∂β1 ∂Z ∂Z †
we notice immediately a huge difference between them: the first one mixes the one-
dimensional bicomplex operators in Z and Z † , while the second one keeps them
separate (!). For example, consider the bicomplex function F (Z) = (Z −Z † )2 ; then
∂F ∂F
(Z) = 0, but (Z) = 2(β1 − β2 ) ,
∂z1 ∂β1
for all Z ∈ BC, thus F is constant with respect to z1 , but not with respect to β1 .
This is tightly related, of course, with the fact that a BC-holomorphic func-
tion when seen as a mapping from C2 to C2 with the cartesian coordinates is a
pair of holomorphic functions which depend on two complex variables and which
have a Cauchy–Riemann type relation, meanwhile the same function written in the
idempotent coordinates becomes a pair of holomorphic functions in one variable
which moreover are independent of each other.
Remark 7.7.4. Let us see now the relation between bicomplex differential operators
and the idempotent representation of hyperbolic numbers. Recalling the formula
∂
for , we regroup the terms in a different way:
∂Z
∂ 1 ∂ ∂ ∂ ∂
= Wφ ◦ −i e+ −i e† ◦ Wφ−1
∂Z 2 ∂1 ∂m1 ∂2 ∂m2
1 ∂ ∂ † ∂ ∂ †
= Wφ ◦ e+ e −i e+ e ◦ Wφ−1 . (7.90)
2 ∂1 ∂2 ∂m1 ∂m2
z1 = x1 + ky2 , z2 = y1 − kx2 ,
7.7. Cartesian versus idempotent representations in BC-holomorphy 177
Going back to formula (7.90), we obtain the following formulation of the bicomplex
differential operator with respect to Z, in terms of the hyperbolic derivatives:
∂ 1 ∂ ∂
= Wφ ◦ −i ◦ Wφ−1 ,
∂Z 2 ∂z1 ∂z2
Ω = Ω1 e + Ω2 e† := β1 e + β2 e† β1 ∈ Ω1 , β2 ∈ Ω2
with Ω1 and Ω2 domains, that is, open and connected sets, in C(i). We will call such
Ω “a product-type domain” in BC. Note that such an Ω should
be seen
as the carte-
Now if, say, G2 ≡ 0 but G1 is not identically zero, then F (Z) = G1 (β1 )e, so
F (Z) = 0 if and only if Z = β1,0 e + λe† , where β1,0 is a zero of G1 and λ is an
arbitrary C(i)-complex number. Therefore the set of zeros of F is a countable (or
finite) union of portions of complex lines parallel to the zero-divisor line Le† =
BCe† ; in this case the complex line that passes through β1,0 e and which is parallel
to BCe† is of the form β1,0 e + BCe† . A similar statement is true for G1 ≡ 0.
Summarizing, we have just proved the following
Theorem 8.1.1. Consider a bicomplex holomorphic function F : Ω → BC which is
not identically zero. Then
1. if F takes at least one invertible value, then its zero set is either the empty
set or a set of isolated points in Ω;
2. if F (Z) ∈ BCe for all Z ∈ Ω, then its zero set is a countable (or finite) union
of portions of complex lines parallel to the zero-divisor line Le† ;
3. if F (Z) ∈ BCe† for all Z ∈ Ω, then its zero set is a countable (or finite)
union of portions of complex lines parallel to the zero-divisor line Le .
Example 8.1.2. Consider the bicomplex holomorphic function F (Z) = Z 2 =
f1 (Z) + jf2 (Z), with f1 (z1 , z2 ) = 2z1 z2 and f2 (z1 , z2 ) = z22 − z12 , which are defined
and holomorphic on C2 (i). Their zero sets are:
V1 = {(z1 , z2 ) ∈ C2 (i) z1 = 0} ∪ {(z1 , z2 ) ∈ C2 (i) z2 = 0} ,
V2 = {(z1 , z2 ) ∈ C2 (i) z1 = z2 } ∪ {(z1 , z2 ) ∈ C2 (i) z1 = −z2 } ,
respectively. Each of them is a union of two complex lines in C2 (i), respectively.
Note that the intersection V1 ∩ V2 = {(0, 0)}, which is just a point and which is
the only zero of F .
A direct computation shows that using the idempotent variables β1 = z1 −iz2
and β2 = z1 + iz2 we get:
F (Z) = β12 e + β22 e† =: G1 (β1 )e + G2 (β2 )e† ,
thus the functions G1 and G2 , which are holomorphic on C(i) with respect to their
corresponding variable, have only one zero: β1 = 0 and β2 = 0, respectively.
Example 8.1.3. Consider now the bicomplex exponential function F (Z) = eZ ,
which is bicomplex holomorphic on BC and takes only invertible values for all
Z ∈ BC. The functions f1 (z1 , z2 ) = ez1 cos z2 and f2 (z1 , z2 ) = ez1 sin z2 have their
zero sets
(2n + 1)π
2
V1 = {(z1 , z2 ) ∈ C (i) cos z2 = 0} = z1 , 2
∈ C (i) n ∈ Z ,
2
V2 = {(z1 , z2 ) ∈ C2 (i) sin z2 = 0} = {(z1 , nπ) ∈ C2 (i) n ∈ Z} ,
which are obviously non-compact sets in C2 (i). But V1 ∩ V2 = ∅, so F has no zeros,
as we already know.
8.2. When bicomplex holomorphic functions reduce to constants 181
for all Z = β1 e+β2 e† ∈ Ω. This is equivalent to G1 (β1 ) = 0 on Ω1 and G2 (β2 ) = 0
on Ω2 . Because Ω1 and Ω2 are domains in the complex plane, this implies that
G1 , G2 are constant functions with respect to the variables β1 , β2 , respectively,
i.e., G1 (β1 ) = a1 ∈ C(i) for all β1 ∈ Ω1 and G2 (β2 ) = a2 ∈ C(i) for all β2 ∈ Ω2 ,
therefore F (Z) = a1 e + a2 e† , a bicomplex constant for all Z ∈ Ω.
Remark 8.2.2. The proof of the theorem above could have been obtained without
using the idempotent representation of F . For example, writing F = f1 + jf2 on
Ω, the existence of F on Ω implies the existence of the complex partial derivatives
of f1 and f2 on Ω with respect to z1 and z2 , related by the Cauchy-Riemann type
∂F
conditions (7.55). Using the fact F (Z) = = 0, we can derive easily that the
∂Z
partial derivatives of f1 and f2 with respect to both z1 and z2 are zero, which
implies that f1 and f2 are constant on the domain Ω, so F is constant on Ω.
Now we are interested in examining under what conditions a bicomplex holo-
morphic function F has zero derivative on a domain Ω.
Theorem 8.2.3. Let F be a bicomplex holomorphic function on a product-type sim-
ply connected domain Ω in BC. Then F is a constant function if and only if any
of the following equivalent conditions holds:
1. Writing F = f11 +if12 +jf21 +kf22 , one of the real functions fk is constant.
In particular, F can be “missing” one (or more) of the real components fk ,
i.e., fk = 0.
2. Writing F = f1 +jf2 , either f1 or f2 is a constant C(i)-function. Similarly for
F = g1 + kg2 . In particular, F can be either C(i)-valued (i.e., f2 = g2 = 0),
or j · C(i)-valued (i.e. f1 = 0), or k · C(i)-valued (i.e., g1 = 0).
3. Writing F = ρ1 + iρ2 , either ρ1 or ρ2 is constant. Similarly for F = γ1 + kγ2 .
In particular, F can be either C(j)-valued (i.e., ρ2 = γ2 = 0), or i·C(j)-valued
(i.e., ρ1 = 0), or k · C(j)-valued (i.e., γ1 = 0).
182 Chapter 8. Some Properties of Bicomplex Holomorphic Functions
respectively Ω2 , domains in C(i), their zeros are isolated, unless they are identically
zero. Thus, it follows that either G1 ≡ 0 on Ω1 or G2 ≡ 0 on Ω2 . In either case,
the result is that F (Z) ∈ S0 for all Z ∈ Ω.
If a = 0, then the following argument applies: the partial derivatives with
respect to both β1 and β2 of the product G1 · G2 are zero:
∂(a2 ) ∂(G1 (β1 ) · G2 (β2 ))
0= = = G1 (β1 ) · G2 (β2 ) ,
∂β1 ∂β1
∂(a2 ) ∂(G1 (β1 ) · G2 (β2 ))
0= = = G1 (β1 ) · G2 (β2 ) ,
∂β2 ∂β2
for all β1 ∈ Ω1 and β2 ∈ Ω2 , so in the case that neither G1 nor G2 are identically
zero, then both G1 and G2 are constant functions on their domains. Therefore, F
is constant. Similarly in the C(j) case.
A special computation is necessary when the hyperbolic modulus of F equals
a hyperbolic constant a = a1 + ka2 , where now a1 , a2 ∈ R. Let F = f1 + if2 , and
write each hyperbolic function f1 = s1 e + t1 e† and f2 = s2 e + t2 e† in idempotent
components (in the intrinsic hyperbolic writing), where s1 , s2 , t1 , t2 are real-valued
functions of the bicomplex variable Z. Then
a21 e + a22 e† = a2 = |F (Z)|2k = (s21 (Z) + s22 (Z))e + (t21 (Z) + t22 (Z))e† ,
therefore
a21 = s21 (Z) + s22 (Z) = G1 (Z) · G1 (Z)
and
a22 = t21 (Z) + t22 (Z) = G2 (Z) · G2 (Z)
for all Z. Because s1 , s2 , t1 and t2 are real-valued functions, then a1 = 0 or a2 = 0
is equivalent to s1 = s2 ≡ 0 or t1 = t2 ≡ 0, respectively, which is equivalent to
F (Z) being a zero-divisor for all Z.
The function G1 (Z) is a complex holomorphic function which depends on β1
only: G1 = G1 (β1 ), thus the function G1 is anti-holomorphic in β1 , so that in the
identity a21 = G1 (β1 ) · G1 (β1 ) we cannot take the derivatives of both sides but we
∂
can apply the Cauchy–Riemann operator which gives:
∂β1
∂(G1 · G1 ) ∂G1
0= (β1 ) = (β1 ) · G1 (β1 ) = G1 (β1 ) · G1 (β1 ) ,
∂β1 ∂β1
∂G1 ∂G1
where we used the fact that = = 0. Since this happens for all Z, so for
∂β1 ∂β 1
all β1 ∈ Ω1 , it follows that either G1 ≡ 0 so G1 ≡ 0, or that G1 is constant. An
8.3. Relations among bicomplex, complex and hyperbolic holomorphies 185
the principal argument Θ = argi (F (Z)) is constant for all Z. Then the expression
ejΘ is an invertible bicomplex number, say, A. Therefore F (Z) = A · f (Z), where
f (Z) = |F (Z)|i is a C(i)-valued function. Because F is bicomplex holomorphic on
Ω, it follows that F is constant. Similarly for the C(j) case.
Corollary 8.2.4. Let F be a bicomplex holomorphic function on a product-type
domain Ω. Each of the following conditions implies that F (Z) ∈ S for all Z ∈ Ω:
1. F (Z) ∈ S for all Z ∈ Ω.
2. If we write F = G1 e + G2 e† , then either G1 or G2 is a constant function.
This happens if and only if F (Z) ∈ S for all Z.
3. If F (Z) = A · f (Z), where A ∈ S and f is either a C(i)-, C(j)-, D-, jC(i)-,
kC(i)-, iC(j)-, kC(j)-, iD- or jD-valued function.
4. |F (Z)|i = 0 for all Z ∈ Ω.
5. |F (Z)|j = 0 for all Z ∈ Ω.
6. |F (Z)|k ∈ S for all Z ∈ Ω.
∂
and anihilate the function F for all Z ∈ Ω, i.e.,
∂Z ∗
∂F ∂F ∂F
(Z) = (Z) = (Z) = 0 . (8.1)
∂Z † ∂Z ∂Z ∗
In this case the bicomplex derivative of F exists for all Z, and it is related to the
fourth differential operator:
∂F
F (Z) = (Z) .
∂Z
∂ ∂
We have seen that the operators and are intimately related, and the
∂Z ∂Z †
∂ ∂
same is true for and .
∂Z ∂Z ∗
One may wonder what happens if instead of having relation (8.1) we impose
one of the following conditions for F for all Z in its domain:
∂F ∂F ∂F
(Z) = (Z) = (Z) = 0 , (8.2)
∂Z ∂Z ∂Z ∗
∂F ∂F ∂F
(Z) = †
(Z) = (Z) = 0 , (8.3)
∂Z ∂Z ∂Z ∗
∂F ∂F ∂F
(Z) = †
(Z) = (Z) = 0 . (8.4)
∂Z ∂Z ∂Z
Equalities (8.2) define what can be called †-anti-holomorphy while equali-
ties (8.3) are about bar-anti-holomorphy and equalities (8.4) are about ∗-anti-
holomorphy. It is clear how to introduce the corresponding derivatives.
We can interpret each of these conditions using all the formulas we developed
so far, in all the possible bicomplex writings. Note that if we compose F with a
change of bicomplex variable W → Z † , then (8.2) is equivalent to
∂F † ∂F † ∂F †
(W ) = (W ) = (W ) = 0 ,
∂W † ∂W ∂W ∗
†
where F † (W ) = (F (W )) . This is equivalent to the fact that the function F † is a
bicomplex holomorphic function of a bicomplex variable W = Z † ; more exactly,
the function F † (Z † ) is simply BC-holomorphic. In the same fashion, F is bicomplex
bar-anti-holomorphic if and only if the function F is bicomplex holomorphic with
respect to Z; F is bicomplex ∗-anti-holomorphic if and only if the function F ∗ is
bicomplex holomorphic with respect to Z ∗ . We leave the details to the reader.
Let us study in more detail the notion of bicomplex †-anti-holomorphy. If we
choose to write F = ρ1 + iρ2 , then (8.2) leads to ρ1 and ρ2 being anti-holomorphic
functions of the two C(j)-complex variables ζ1 and ζ2 , i.e., they are holomorphic
188 Chapter 8. Some Properties of Bicomplex Holomorphic Functions
with respect to the j-conjugate variables ζ1∗ and ζ2∗ , related by the Cauchy-Riemann
type system:
∂ρ1 ∂ρ2 ∂ρ1 ∂ρ2
∗ = ∗, ∗ =− ∗.
∂ζ1 ∂ζ2 ∂ζ2 ∂ζ1
For the same condition (8.2), using the idempotent representation F = Γ1 e+Γ2 e† ,
where Z = γ1 e + γ2 e† with γ1 , γ2 ∈ C(j), we obtain that Γ1 and Γ2 are complex
anti-holomorphic functions depending on only one variable: γ1∗ for Γ1 and γ2∗ for
Γ2 , respectively. Similar statements are true for the other two types of bicomplex
anti-holomorphy.
In the C(i)-idempotent representation F = G1 e+G2 e† , where Z = β1 e+β2 e†
with β1 , β2 ∈ C(i), equation (8.2) leads to the conclusion that G1 and G2 are
complex holomorphic in variables β2 and β1 , respectively: note here the reversed
role of the complex variables β1 and β2 .
and
argD f (Z0 ) = θ1 e + θ2 e† ∈ D.
Next, take a smooth hyperbolic curve Γ passing through Z0 and let T be the
hyperbolic tangent line to Γ at Z0 . By the previous Lemma, f (Γ) = Λ = λ1 e+λ2 e†
is also a smooth hyperbolic curve with hyperbolic tangent line P at W0 = f (Z0 ).
Since Γ = γ1 e + γ2 e† has a parametrization
where ϕ1 and ϕ2 are, respectively, the parametrizations of the smooth, real curves
γ1 and γ2 in C(i), then one can take the projections of T and P on BCe and BCe†
and write:
T = t 1 e + t 2 e† , P = p 1 e + p 2 e† .
Writing also Z0 = β10 e + β2 e† = ϕ1 (u0 )e + ϕ2 (v0 )e† , we note that t1 is the (real)
tangent line to γ1 at ϕ1 (u0 ) as well as t2 , p1 , p2 are the (real) tangent lines to γ2 ,
λ1 λ2 at ϕ2 (v0 ), F1 (ϕ1 (u0 )), F2 (ϕ2 (v0 )) respectively.
Denote now by μ1 and μ2 the real numbers such that tan μ1 and tan μ2 are
the (real) slopes of t1 and t2 respectively. Similarly, the real numbers κ1 and κ2
are such that tan κ1 and tan κ2 are the (real) slopes of p1 and p2 . Hence, the
hyperbolic slopes of the hyperbolic lines T and P are, respectively,
But θ1 = arg F1 (β10 ) and θ2 = arg F2 (β2 (β20 ) hence
κ1 = μ 1 + θ 1 and κ2 = μ 2 + θ 2
" passing
We are ready to complete the proof. Take now two curves Γ and Γ
through Z0 , then for each of the (8.5) it holds that:
κ = μ + argD f (Z0 ),
" + argD f (Z0 ).
"=μ
κ
Substracting them gives:
κ−κ
"=μ−μ
"
which means that f is a D-conformal mapping ar Z0 .
Remark 8.5.4. Consider also an interpretation of the hyperbolic modulus of f (Z0 ).
By the properties of such moduli, one has:
f (Z0 ) = lim |f (Z) − f (Z0 )|D
D Z→Z0 |Z − Z0 |D
for Z − Z0 ∈ S0 . Hence, for any Z close to Z0 , Z − Z0 ∈ S0 , with respect to the
hyperbolic modulus there holds:
f (Z) − f (Z0 ) ≈ f (Z0 ) · Z − Z0 . (8.6)
D D D
In analogy with the case of complex holomorphic functions, we say that a bicom-
plex holomorphic function with the derivative at Z0 different from zero and zero-
divisors,
realizes locally, at Z0 , a hyperbolic homothety, with the hyperbolic coeffi-
cient f (Z0 )D , in all directions except those of the cone of zero-divisors generated
by Z0 .
Proof. Since Ω = Ω1 e+Ω2 e† we can apply the complex Riemann mapping theorem
to Ω1 and Ω2 obtaining the bijective complex conformal mappings F1 and F2
onto the unit disk B(0, 1) ⊂ C(i), which are complex holomorphic functions with
derivatives different from zero. Set
F := F1 e + F2 e† .
then
∂ ∂ ∂ ∂ 1
◦ = ◦ = ΔR2 . (9.1)
∂z ∂z ∂z ∂z 4
Similarly, if ΔRk denotes the Laplace operator in Rk , DCR denotes the Cauchy-
Riemann operator of Clifford analysis in Rn+1 , and DDir denotes the Dirac oper-
ator in Rn (see [23] and [9]), then
Factorization (9.1) manifests the essence of the relation between complex holomor-
phic functions and harmonic functions. First of all, note that the Laplace operator
ΔR2 acts initially on real-valued functions but its action extends onto complex
valued functions component-wise: if f = u + iv, then
Thus equality (9.1) holds on complex-valued functions (of class C 2 , not just C 1 ).
∂f
Let f = u + iv be a holomorphic function, that is, = 0, hence by (9.1)
∂z
∂ ∂f
ΔR2 [f ] = 4 = 0 and thus f is a harmonic function. Reciprocally, taking
∂z ∂z
∂u
a real-valued harmonic function u, consider (which is a formal derivative,
∂z
that is, the result of the action of an operator, not the “honest” derivative of a
holomorphic function), one has:
+ ,
∂ ∂u 1
= ΔR2 [u] = 0 ,
∂z ∂z 4
∂u
which means that the complex-valued function generated by the harmonic
∂z
function u is holomorphic.
∂2 ∂2
ΔC2 (i) := + ; (9.2)
∂z12 ∂z22
∂2 ∂2
ΔC2 (j) := 2 + 2 ; (9.3)
∂ζ1 ∂ζ2
∂2 ∂2
ΔD := 2 + 2 . (9.4)
∂z1 ∂z2
Operators (9.2) and (9.3) are called complex (C(i) and C(j) respectively) Lapla-
cians and (9.4) is the hyperbolic Laplacian. The first of them acts on C(i)-valued
9.2. Complex and hyperbolic Laplacians 195
holomorphic functions of two complex variables z1 and z2 ; the second acts on C(j)-
valued holomorphic functions of the complex variables ζ1 and ζ2 ; and the third
acts on D-valued holomorphic functions of the hyperbolic variables z1 and z2 .
∂
Although we wrote in (9.2)–(9.4) the formal partial derivatives , etc.,
∂z1
in fact we mean authentic partial derivatives with respect to the corresponding
variables z1 , z2 , ζ1 , ζ2 , z1 , z2 . Of course in this situation the formal and the authen-
tic derivatives coincide but at the same time a confusion may arise, so that we
rewrite (9.2)–(9.4) as
ΔC2 (i) = ∂z2 + ∂z2 ; (9.5)
1 2
This is in keeping with the notation we adopted where we wrote complex partial
derivatives as fz 1 , etc. Hence ∂z2 f = fz1 , etc.
1
The next step consists in extending the operators (9.5)–(9.7) onto bicom-
plex-valued functions. It turns out that how we write bicomplex numbers becomes
important and it depends on the operators in (9.5)–(9.7). Indeed, for the operator
ΔC2 (i) the bicomplex function F should be holomorphic in the sense of two complex
variables and if we want the same for its components, then we are forced to consider
F as F = f1 + jf2 . Now we set:
ΔC2 (i) [F ] := ΔC2 (i) [f1 ] + jΔC2 (i) [f2 ] .
A similar reasoning holds for the other two operators: for ΔC2 (j) we take F =
ρ1 + iρ2 and we set:
ΔC2 (j) [F ] := ΔC2 (j) [ρ1 ] + iΔC2 (j) [ρ2 ] ;
for ΔD2 we take F = f1 + if2 and we set:
ΔD2 [F ] := ΔD2 [f1 ] + iΔD2 [f2 ] .
The analogues of the formula (9.1) arise if one uses the corresponding opera-
tors in formulas (7.49), (7.50), (7.51) and (7.52); more exactly the operators should
∂
be taken in an appropriate form. For instance, if we want to use the operators
∂Z †
∂
and in the form
∂Z
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
= + j , = − j ,
∂Z † 2 ∂z1 ∂z2 ∂Z 2 ∂z1 ∂z2
then we are forced to write the variable as Z = z1 + jz2 and the function F as
F = f1 + jf2 . With this understanding we have that
∂2 ∂2 ∂ ∂
ΔC2 (i) = 2 + 2 =4 , (9.8)
∂z1 ∂z2 ∂Z ∂Z †
196 Chapter 9. Second Order Complex and Hyperbolic Differential Operators
∂f1
which means that the BC-valued function generated by the complex harmonic
∂Z
function f1 is BC-holomorphic.
The operators in (9.2) and (9.3) are dealt with in exactly the same way,
although now other first-order operators enter into the game. In the case of Z =
ζ1 + jζ2 we take
∂ 1 ∂ ∂ ∂ 1 ∂ ∂
= −i , = +i ,
∂Z 2 ∂ζ1 ∂ζ2 ∂Z 2 ∂ζ1 ∂ζ2
arriving at
∂2 ∂2 ∂ ∂
ΔC2 (j) = 2 + 2 =4 ; (9.9)
∂ζ1 ∂ζ2 ∂Z ∂Z
arriving at
∂2 ∂2 ∂ ∂
ΔD = 2 + 2 =4 . (9.10)
∂z1 ∂z2 ∂Z ∂Z ∗
9.3. Complex and hyperbolic wave operators 197
So, we have provided each of the operators (9.2)-(9.4) with an adequate func-
tion theory. There is a fine point here: it is, in a sense, one and the same theory of
BC-holomorphic functions, but where different aspects of the latter are taken into
account. This “common” function theory allows us to realize the similarities and
the differences between the three operators (9.2)-(9.4). Indeed, (9.2) and (9.3) look
identical from the viewpoint of classical complex analysis, but viewing them from
the bicomplex perspective reveals subtle differences between them. At the same
time, (9.4) seems to be quite different with any of (9.2) and (9.3) but bicomplex
functions, again, show that there exists a deep underlying unity between all three
of them.
The direct relations between the null solutions of any of the operators ΔC2 (j)
and ΔD2 and the BC-holomorphic function theory is established in the same way
as we did for the operator ΔC2 (i) .
turn the operator C2 (i) into ΔC2 (i) and the operator C2 (j) into ΔC2 (j) . But they
can be factorized directly:
∂ ∂ ∂ ∂
C2 (i) = +k · −k
∂w1 ∂w2 ∂w1 ∂w2
∂ ∂
=4 ;
∂Z ∂Z †
compare with (9.8). Although formally they are the same operators, they are
employed in other forms taken in the table in Section 4. In the same fashion,
∂ ∂ ∂ ∂
C2 (j) = +k · −k
∂ω1 ∂ω2 ∂ω1 ∂ω2
∂ ∂
=4 ,
∂Z ∂Z
198 Chapter 9. Second Order Complex and Hyperbolic Differential Operators
where we take, again, another form of writing the operators involved. Somewhat
paradoxically, the same bicomplex “tricks” do not work for the “hyperbolic wave
operator”
∂2 ∂2
2 −
∂w1 ∂w22
which can neither be factorized directly nor reduced to (9.4); this is because in
the hyperbolic world there is only one hyperbolic-type imaginary unit and there
are no complex-type imaginary units.
One can consider another hyperbolic Laplacian:
∂2 ∂2
+
∂w21 ∂w22
reduces it to (9.4).
thus the series of bicomplex numbers (10.1) is convergent if and only if both
complex series in (10.2) are convergent.
∞
$
A typical example is the bicomplex geometric series Zn = 1 + Z + Z2 +
n=1
· · · + Z n + · · · which converges for Z in the bicomplex unitary ball, that is, the
bicomplex ball of hyperbolic radius 1, and which diverges outside this ball; this
is because the bicomplex geometric series is equivalent to the pair of complex
geometric series:
$∞ ∞
$ ∞
$
Zn = β1n e + β2n e† .
n=1 n=1 n=1
∞
$
Theorem 10.1.1 (Cauchy criteria). The series Zn converges if and only if the
n=1
sequence of its partial sums is a D-Cauchy sequence, i.e., for any positive hyperbolic
Theorem 10.2.1. If all the functions fn are continuous on Ω and the sequence
converges D-uniformly on Ω, then the limit function f is continuous on Ω.
Proof. Mimics the complex functions case using the properties of hyperbolic pos-
itive numbers instead of real positive numbers.
Proof. Since each fn is bicomplex holomorphic, then fn (Z) = F1,n (β1 )e+F2,n (β2 )e†
where for any n the functions F1,n are holomorphic in Ω1 and all F2,n are holo-
morphic in Ω2 . If K is a compact subset in Ω and K = K1 e + K2 e† , then K1 is
∞
a compact subset in Ω1 and K2 is a compact subset in Ω2 . Moreovoer, {fn }n=1
∞
converges D-uniformly on K if and only if {F1,n }n=1 converges uniformly on K1
∞
and {F2,n }n=1 converges uniformly on K2 .
Note that a priori the limit function f does not need to be bicomplex holo-
morphic, in general, thus, in its idempotent representation
the components F1 and F2 would depend on both variables β1 and β2 . But the
∞
complex Weierstrass theorem implies that the sequence {F1,n }n=1 converges to
∞
F1 uniformly on compacts and {F2,n }n=1 converges to F2 uniformly on compacts,
204 Chapter 10. Sequences and Series of Bicomplex Functions
where all the functions fn are defined on the set Ω in BC. Their behavior is char-
$n
acterized by the sequence Sn (Z) = fk (Z) of partial sums, hence the properties
k=1
of sequences (convergence at a point, D-uniform convergence on the set, etc.) ap-
∞
ply to the sequence {Sn (Z)}n=1 and become the respective properties of series
of functions. For instance, in order that the series of functions (10.3) converges
D-uniformly on Ω it is necessary and sufficient that for any positive hyperbolic
number ε there exists N = N (ε) such that fN +1 (Z) + · · · + fN +m (Z)D ≺ ε for
any Z ∈ Ω and any m ∈ N (Cauchy’s criteria).
∞
$ ∞
$
A series fn (Z) is called D-absolutely convergent if the series fn (Z)
D
n=1 n=1
converges on Ω.
Theorem 10.3.1 (Weierstrass test for D-uniform convergence). Given a series of
$∞ $∞
functions fn (Z), if there exists a convergent series An of hyperbolic non-
n=1 n=1
negative numbers such that the inequality fn (Z) An holds for n = n∗ , n∗ +
$∞
1, . . . , for some n∗ ∈ N and for all Z in Ω, then the series fn (Z) converges
n=1
D-uniformly and D-absolutely on Ω.
Proof. Consists in applying Cauchy’s criteria twice.
The properties of sequences of functions give rise to the following statements
about series of functions.
• If the functions fn are continuous on Ω and the series converges D-uniformly
on Ω, then the sum S(Z) is a continuous function.
10.4. Bicomplex power series 205
∞
$
• (Weierstrass theorem.) If every fn is bicomplex holomorphic in Ω and fn (Z)
n=1
converges D-uniformly on compact subsets of Ω, then the sum S is bicom-
∞
$
plex holomorphic in Ω and the series fn(k) (Z) converges D-uniformly on
n=1
compact subsets in Ω to S (k) (Z) for any k ∈ N.
Let Z" ∈ S be a
point of convergence.
Take K to be a closed subset of Bρ
Z
and consider the set Z ∈ K . It is D-bounded and, hence, it has a D-
Z" D
Z
supremum, supD ; what is more, supD Z ∈ K =: q ≺ 1. Since the se-
Z" D
∞
$
ries An Z"n converges, then lim An Z" n = 0, thus the sequence An Z" n is D-
n→∞
n=0
bounded: there exists a positive hyperbolic number M such that An Z"n D M
for n ∈ N. For Z ∈ K we get now an estimate
n
An Z n = An Z" n · Z M q n .
D
Z" D
The series with general term M q n converges and it may serve as a majorant for
∞
$
the series An Z n on K; the Weierstrass theorem allows us to conclude that
n=0
∞
$
An Z n
converges D-absolutely and D-uniformly on K. Since any point of Bρ
n=0
belongs to some subset K, the theorem is proved.
As was noted above, if ρ is a zero-divisor, i.e., ρ = ρ1 e or ρ = ρ2 e† , then
all the points of Sρ are zero-divisors of the same form. What can be said if
$∞
Z" ∈ Sρ and the series An Z"n converges? Writing An = an e + bn e† we get:
n=0 ∞ ∞
$∞ $∞
$ $
An Z"n = an e + bn e† β"1 e + β"2 e† = an β"n e +
1 bn β"n e† .
2
n=0 n=0 n=0 n=0
Since Z" is of one of the forms: Z" = β"1 e or Z" = β"2 e† , then we can make
$∞
a conclusion for one of the bicomplex components of the series An Z n =
∞ ∞ n=0
∞
$ $ $
an β1n e + bn β2n e† . Indeed, if ρ = ρ1 e, then the series an β1n e
n=0 n=0 n=0
∞
$
converges for |β1 | < ρ1 but we have no information for the series bn β2n e† ;
n=0
similarly for ρ = ρ2 e† . Of course, this does not mean that there is no information
in principle; we have just taken an “unlucky” point Z. "
Several possibilities arise now. Let, first, one of the idempotent series converge at
the origin only, then the other may converge in a disk of finite radius which gives
the case (4) or in the whole C(i) which gives the case (3); illustrations for (3) are
∞
$
$∞
1 n † 1 n n †
n! β1n e + β e or β e + n! β2 e
n=0
n! 2 n=0
n! 1
∞
$
where γn z n is a complex series with a (real) radius of convergence r > 0. Sec-
n=0
ond, both idempotent series converge on a disk which may be of a finite or infinite
∞
$
radius thus justifying the cases (5) and (6); for the same complex series γn z n
∞
∞
n=0
$ 1 n † $ 1 n n †
any of the bicomplex series γn β1n e + β e or β e + γn β2 e
n=0
n! 2 n=0
n! 1
illustrates the former case.
Analogously to the case of complex analysis, we may speak in all the cases
about bicomplex balls BR of hyperbolic radii R where:
208 Chapter 10. Sequences and Series of Bicomplex Functions
S (n) (Z0 )
Proof. By Proposition 10.4.4, An = , thus
n!
∞
$ $∞
n S (n) (Z0 ) n
S(Z) = An (Z − Z0 ) = . (Z − Z0 )
n=0 n=0
n!
Then the Taylor series converges in the bicomplex ball Bd (Z0 ) where its sum co-
incides with f (Z):
$∞
f (n) (Z0 ) n
f (Z) = (Z − Z0 )
n=0
n!
Z0 = β10 e + β20 e† ,
where β10 ∈ Ω1 , β20 ∈ Ω2 . Denote by d1 the Euclidean distance between β10 and
∂Ω1 , and by d2 the Euclidean distance between β20 and ∂Ω2 , then by the complex
Taylor theorem one has:
$∞ (n)
F1 (β10 ) n
F1 (β1 ) = β1 − β10
n=0
n!
for β1 ∈ B β10 , d1 , and
$∞ (n)
F2 (β20 ) n
F2 (β2 ) = β2 − β20
n=0
n!
for β2 ∈ B β20 , d2 . Hence
210 Chapter 10. Sequences and Series of Bicomplex Functions
In addition to the book of Price [56], we refer the reader also to publications
such as [40, 41, 84], where certain properties of bicomplex sequences and series are
studied.
Chapter 11
∂F ∂F ∂F ∂F
dF = dZ + dZ + dZ † + dZ ∗ .
∂Z ∂Z ∂Z † ∂Z ∗
Note that this is just a “bicomplex combination” of the real differentials of the
components of F ; hence the function itself and the variable Z can be written in
any form as well as all the differentials and the differential operators; for instance,
dZ = dx1 + idy1 + jdx2 + kdy2 = dz1 + jdz2 = dζ1 + idζ2 = . . ., dZ = dx1 −
idy1 + jdx2 − kdy2 = dz 1 + jdz 2 = . . ..
Next, consider the action of the exterior differentiation operator on the dif-
ferential form F dZ; we have that
d(F dZ) = dF ∧ dZ
∂F ∂F ∂F † ∂F ∗
= dZ + dZ + dZ + dZ ∧ dZ
∂Z ∂Z ∂Z † ∂Z ∗ (11.1)
∂F ∂F ∂F
= dZ ∧ dZ + †
dZ † ∧ dZ + dZ ∗ ∧ dZ.
∂Z ∂Z ∂Z ∗
The same can be done with the three other differential forms, that is, F dZ, F dZ † ,
F dZ ∗ , arriving at
∂F ∂F ∂F
d(F dZ) = dZ ∧ dZ + dZ † ∧ dZ + dZ ∗ ∧ dZ; (11.2)
∂Z ∂Z † ∂Z ∗
∂F ∂F ∂F
d(F dZ † ) = dZ ∧ dZ † + dZ ∧ dZ † + ∗
dZ ∗ ∧ dZ † ; (11.3)
∂Z ∂Z ∂Z
∂F ∂F ∂F
d(F dZ ∗ ) = dZ ∧ dZ ∗ + dZ ∧ dZ ∗ + dZ † ∧ dZ ∗ . (11.4)
∂Z ∂Z ∂Z †
Of course, each of the formulas (11.2), (11.3) and (11.4) can be obtained also by
making the corresponding change of variable.
Next, take Γ to be a two-dimensional, piecewise smooth, oriented surface in
Ω whose boundary γ = ∂Γ is a piecewise smooth curve.
Then integrating both parts of (11.1) and applying Stokes theorem we get:
/ /
F dZ = d(F dZ)
γ Γ
/
(11.5)
∂F ∂F † ∂F ∗
= dZ ∧ dZ + dZ ∧ dZ + dZ ∧ dZ .
Γ ∂Z ∂Z † ∂Z ∗
Similar formulas arise if one uses (11.2), (11.3) and (11.4).
Formula (11.5) contains several special cases which are of interest by them-
selves. In particular, we can take F as
F (z1 , z2 ) = f1 (z1 , z2 ) + jf2 (z1 , z2 )
where f2 (z1 , z2 ) = 0 in Ω and f1 is holomorphic in the sense of two complex
variables z1 and z2 which leads to
/ /
∂f1 ∂f1
f1 (z1 , z2 )(dz1 + jdz2 ) = +j j dz1 ∧ dz2
γ Γ ∂z1 ∂z2
∂f1 ∂f1
since = = 0 in Ω and
∂Z ∂Z ∗
dZ † ∧ dZ = 2 j dz1 ∧ dz2 .
Separation of the complex components results in the formulas
/ /
∂f1
f1 (z1 , z2 )dz1 = − dz1 ∧ dz2 ;
γ Γ ∂z2
11.1. Stokes’ formula compatible with the bicomplex CR operators 213
/ /
∂f1
f1 (z1 , z2 )dz2 = dz1 ∧ dz2 .
γ Γ ∂z1
The formulas remain true if, in addition, f does not depend on z2 :
/
f1 (z1 )dz1 = 0,
γ
/ /
f1 (z1 )dz2 = f1 (z1 )dz1 ∧ dz2 .
γ Γ
Analogously, if f1 does not depend on z1 , then
/ /
f1 (z2 )dz1 = − f1 (z2 )dz1 ∧ dz2 ,
γ Γ
/
f1 (z2 )dz2 = 0.
γ
But F can be taken also as
F (ζ1 , ζ2 ) = ρ1 (ζ1 , ζ2 ) + iρ2 (ζ1 , ζ2 )
where ρ2 (ζ1 , ζ2 ) = 0 in Ω and ρ1 is holomorphic in the sense of two complex
variables ζ1 and ζ2 ; now we can repeat the reasoning and arrive at very similar
formulas and conclusions.
Finally, F can be taken as
F (z1 , z2 ) = f1 (z1 , z2 ) + if2 (z1 , z2 )
where z1 and z2 are hyperbolic variables, f1 and f2 are D-valued functions; what is
more, we assume that f2 (z1 , z2 ) = 0 in Ω and that f1 is holomorphic in the sense
∂f1 ∂f1
of two hyperbolic variables z1 and z2 . Then = = 0 and dZ ∗ ∧ dZ =
∂Z ∂Z †
2 idz1 ∧ dz2 ; thus (11.5) gives the equality
/ /
∂f1 ∂f1
f1 (z1 , z2 )(dz1 + idz2 ) = +i i dz1 ∧ dz2 .
γ Γ ∂z1 ∂z2
Separation of the hyperbolic components gives:
/ /
∂f1
f1 (z1 , z2 )dz1 = − dz1 ∧ dz2 ,
γ Γ ∂z 2
/ /
∂f1
f1 (z1 , z2 )dz2 = dz1 ∧ dz2 .
γ Γ ∂z1
The same comments as above can be made although the situation is rather different
because holomorphic functions of two hyperbolic variables do not have yet their
theory.
But, of course, for us the most important consequence of (11.5) is the bicom-
plex Cauchy integral theorem.
214 Chapter 11. Integral Formulas and Theorems
∂F ∂F ∂F
Proof. Since F is bicomplex holomorphic, then = = = 0 in Ω, and
∂Z ∂Z † ∂Z ∗
the result follows from (11.5).
/ / b
g(Z) dZ = g(ϕ(t)) ϕ (t) dt
γ a
/ b
= g1 (ϕ1 (t))e + g2 (ϕ2 (t))e† ϕ1 (t)e + ϕ2 (t)e† dt
a
/ b
= g1 (ϕ1 (t)) ϕ1 (t)e + g2 (ϕ2 (t)) ϕ2 (t)e† dt
a
/ /
†
=e g1 (β1 ) dβ1 + e g2 (β2 ) dβ2 ,
γ1 γ2
that is, / / /
g(Z) dZ = e g1 (β1 ) dβ1 + e† g2 (β2 ) dβ2 . (11.6)
γ γ1 γ2
As a matter of fact, the first equality is the definition of the integral of the
bicomplex differential form g(Z) dZ along γ, and it is easy to see that this definition
is well posed.
A similar reasoning applies to the surface Γ and the differential form g(Z) dZ∧
dZ ∗ . Indeed,
/ /
∗
g(Z) dZ ∧ dZ = g(Z) dβ1 e + dβ2 e† ∧ dβ 1 e + dβ 2 e†
Γ
/Γ
= g1 (β1 )e + g2 (β2 )e† dβ1 ∧ dβ 1 e + dβ2 ∧ dβ 2 e†
/Γ
= g1 (β1 ) dβ1 ∧ dβ 1 e + g2 (β2 ) dβ2 ∧ dβ 2 e†
Γ
/
∂ψ1 ∂ψ 1 ∂ψ1 ∂ψ 1
= g1 (ψ1 (u, v)) · − · e
(u,v) ∂u ∂v ∂v ∂u
∂ψ2 ∂ψ 2 ∂ψ2 ∂ψ 2
+ g2 (ψ2 (u, v)) · − · e† du dv
∂u ∂v ∂v ∂u
/ /
= g1 (β1 ) dβ1 ∧ dβ 1 e + g2 (β2 ) dβ2 ∧ dβ 2 e† ,
Γ1 Γ2
that is,
/ /
/
∗
g(Z) dZ ∧dZ = g1 (β1 ) dβ1 ∧ dβ 1 e+ g2 (β2 ) dβ2 ∧ dβ 2 e† . (11.7)
Γ Γ1 Γ2
Remark 11.2.1. Under the same hypotheses, we say that the function g is integrable
in the improper sense along Γ if the function g is integrable in the improper sense
along the domain Γ for = 1 and = 2.
Theorem 11.2.2 (Bicomplex Borel–Pompeiu formula). Let g ∈ C 1 (Ω) be such that
g(Z) = g1 (β1 )e + g2 (β2 )e† , Z = β1 e + β2 e† , and let γ and Γ be as described above.
216 Chapter 11. Integral Formulas and Theorems
/ / ∂g
1 g(t) dt 1 ∂t∗ dt ∧ dt∗ ,
g(Z) = + (11.8)
2πi γ t−Z 2πi Γ t−Z
∂ ∂ ∂
where t = t1 e + t2 e† and =e + e† .
∂t∗ ∂t1 ∂t2
Proof. One has:
g(Z) = g1 (β1 )e + g2 (β2 )e† .
Using the complex Borel–Pompeiu formula gives:
∂g1
/ /
1 g1 (t1 ) dt1 1 ∂t1
g(Z) = e+ dt1 ∧ dt1 e
2πi γ1 t1 − β 1 2πi Γ1 1 − β 1
t
∂g2
/ /
1 g2 (t2 ) dt2 † 1 ∂t2
+ e + dt2 ∧ dt2 e†
2πi γ2 t2 − β 2 2πi Γ2 t2 − β2
Note that we used in the proof formulas (11.6) and (11.7) and Remark 11.2.1.
Theorem 11.2.3 (The bicomplex Cauchy integral representation). Let Ω be a
product-type domain in BC, let f be a bicomplex holomorphic function in Ω, and
let Z be an arbitrary point in Ω. Then, for any surface Γ ⊂ Ω passing through Z
and with the above described properties the Cauchy representation formula holds:
/
f (t)
F (Z) = dt,
γ t−Z
where γ = ∂Γ.
11.2. Bicomplex Borel–Pompeiu formula 217
[26] L.D. Egorova, L.I. Krjuckova, L.B. Lobanova. Bicomplex and bidual
spaces. Moskov. Oblast. Ped. Inst. Ucen. Zap. v. 262 (1969), 76–103.
[27] R. Fueter. Analytische Funktionen einer Quaternionen Variablen. Comm.
Math. Helv. 4 (1932), 9-20.
[28] S.G. Gal. Introduction to geometric function theory of hypercomplex vari-
ables. Nova Science Publishers, Inc. Chapter 4 (2004), xvi + 319.
[29] V. Garant-Pelletier, D. Rochon. On a generalized Fatou-Julia theorem
in multicomplex spaces. Fractals v. 17, No. 3 (2009), 241–255.
[30] G. Gentili, D.C. Struppa. A new theory of regular functions of a quater-
nionic variable. Adv. Math. v. 216, No. 1 (2007), 279-301.
[31] G. Gentili, C. Stoppato, D.C. Struppa. Regular Functions of a Quater-
nionic Variable. Springer Verlag (2013).
[32] R. Gervais-Lavoie, L. Marchildon, D. Rochon. Infinite-dimensional
bicomplex Hilbert spaces. Ann. Funct. Anal. v. 1 No. 2 (2010), 75–91.
[33] R. Gervais-Lavoie, L. Marchildon, D. Rochon. The bicomplex quan-
tum harmonic oscillator. Nuovo Cimento Soc. Ital. Fis. B, v. 125, No. 10
(2010), 1173–1192.
[34] R. Gervais-Lavoie, L. Marchildon, D. Rochon. Finite-dimensional
bicomplex Hilbert spaces. Adv. Appl. Clifford Algebr. v. 21, No. 3 (2011),
561–581.
[35] K. Gürlebeck, F. Kippig. Complex Clifford-Analysis and Elliptic Bound-
ary Problems. Adv. Appl. Clifford Algebras, v. 5, No. 1 (1995), 51–62.
[36] W.R. Hamilton. On quaternions, or on a new system of imaginaries in
algebra. Philosophical Magazine. Vol. 25, no. 3 (1844), 489–495.
[37] W.R. Hamilton. Lectures on Quaternions: Containing a Systematic State-
ment of a New Mathematical Method. Dublin: Hodges and Smith (1853).
[38] H. Kabadayi, Y. Yayli. Homothetic Motions at E4 with Bicomplex Num-
bers. Adv. Appl. Clifford Algebras, v. 21, No. 2 (2002), 541–546.
[39] Krantz. Several Complex Variables. Second Edition, AMS Chelsea Pub-
lishing (2001).
[40] R.S. Krausshar. Eisenstein Series in Complexified Clifford Analysis.
Comp. Methods and Func. Theory, v. 2, No. 1 (2002), 29–65.
[41] J. Kumar, R.K. Srivastava. On a class of entire bicomplex sequences.
South East Asian J. Math. Math. Sci. v. 5, No. 3 (2007), 47–67.
[42] J. Kumar, R.K. Srivastava. A note on poles of the bicomplex Riemann
zeta function. South East Asian J. Math. Math. Sci. v. 9, No. 1 (2010),
65–75.
222 Bibliography
[90] R.K. Srivastava. Certain points in the theory of bicomplex numbers. Math.
Student, v. 70, No. 1–4 (2001), 153–160.
[91] R.K. Srivastava. Bicomplex numbers: analysis and applications. Math.
Student, v. 72, No. 1–4 (2003), 69–87.
[92] R.K. Srivastava. Certain topological aspects of bicomplex space. Bull. Pure
Appl. Math.v. 2, No. 2 (2008), 222–234.
[93] R.K Srivastava, S. Singh. Certain bicomplex dictionary order topologies.
Int. J. Math. Sci. Eng. Appl. v. 4, No. 3 (2010), 245–258.
[94] N.T. Stelmasuk. On certain linear partial differential equations in dual
and bicomplex algebras. An. Sti. Univ. ”Al. I. Cuza” Iasi Sect. I (N.S.), v. 9
(1963), 63–72.
[95] N.T. Stelmasuk. Some linear partial differential equations in dual and
bicomplex algebras. Izv. Vyss. Ucebn. Zaved. Matematika, v. 40, No. 2 (1964),
136–142.
[96] D.C. Struppa, A. Vajiac, M.B. Vajiac. Remarks on Holomorphicity
in Three Settings: Complex, Quaternionic, and Bicomplex. Hypercomplex
Analysis and Applications, Trends in Mathematics, Springer, I. Sabadini, F.
Sommen editors (2011), 261–274.
[97] A. Sudbery Quaternionic Analysis. Mathematical Proceedings of the Cam-
bridge Philosophical Society v. 85 (1979), 199?225.
[98] V.A. Tret’jakov. Some properties of mappings in the space accompaning
the algebra of bicomplex numbers in modulus (Russian). Application of func-
tional analysis in the approximation theory (Russian), Kalinin. Gos. Univ.,
Kalinin (1979), 129–136.
[99] V.A. Tret’jakov. On the properties of some elementary functions that
are defined on the algebra of bicomplex numbers (Russian). Mathematical
analysis and the theory of functions (Russian), Moskov. Oblast. Ped. Inst.,
Moscow (1980), 99–106.
[100] I.M. Yaglom. Complex Numbers in Geometry. Academic Press, New York-
London (1968), 243 pp.
[101] K. Yoneda. An integration theory in the general bicomplex function theory.
Yokohama Math. J. v. 1 (1953), 225–262.
[102] A. Zireh. A generalized Mandelbrot set of polynomials of type Ed for bi-
complex numbers. Georgian Math. J. v. 15, No. 1 (2008), 189–194.
[103] V.A. Zorich. Mathematical Analysis, Volumes I and II. Springer–Verlag
Berlin Heidelberg (2004).
Index
Real differentiability
Idempotent
In C( j)-complex terms, 154
Elements, 15
In C(i)-complex terms, 153
Representation, 15
In bicomplex terms, 156
Idempotent real variables, 162
In hyperbolic terms, 155
Invertible bicomplex numbers
Real exponential function, 119
Characterization, 14
Real idempotent components and real
Idempotent representation, 19
cartesian components
Jacobi matrix Matrix of change of variable, 167
Relation with bicomplex deriv- Relation between them, 167
ability, 145, 146 Real Laplace operator
Definition, 193
Moduli of bicomplex numbers Factorization, 193
Definitions, 9 in n variables, 193
Idempotent representation, 18 Real lines in BC, 77
Properties, 11 Real partial derivatives of a bicom-
plex function
Negative hyperbolic numbers, 7 Definition, 136
Non-negative hyperbolic numbers Its relation with bicomplex deriv-
cartesian representation, 7 ability, 144
Idempotent representation, 23 Real polynomials
Non-positive hyperbolic numbers, 7 Definition, 113
Idempotent representation, 23 Properties, 114
Real straight lines
Operator of the change of variable Complex form, 73
and BC-modules, 168 in R2 , 73
and gradient, 170 Parametric representation, 76
and operators acting on BC-modules, Riemann Mapping Theorem
168 Bicomplex case, 190
and real partial derivatives, 169 Complex case, 190
Definition, 168
Sequences of bicomplex functions
Partial order on D D-uniformly convergent
Definition, 41 Bicomplex analogue of Weier-
Properties, 42 strass’ Theorem, 203
Index 231