You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/345990039

Clifford Analysis with Indefinite Signature

Preprint · November 2020

CITATIONS READS
0 12

2 authors, including:

Matvei Libine
Indiana University Bloomington
31 PUBLICATIONS   223 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Matvei Libine on 22 February 2021.

The user has requested enhancement of the downloaded file.


Clifford Analysis with Indefinite Signature
Matvei Libine∗and Ely Sandine†
arXiv:2011.08289v1 [math.CV] 16 Nov 2020

November 18, 2020

Abstract
We extend constructions of classical Clifford analysis to the case of indefinite non-
degenerate quadratic forms. We define (p, q)-left- and right-monogenic functions by
means of Dirac operators that factor a certain wave operator. We prove two different
versions of Cauchy’s integral formulas for these functions. The two formulas arise from
dealing with singularities in distinct ways, and are inspired by the methods of [L, FL].
These results indicate the merit of these methods for dealing with singularities.

1 Introduction
Many results of complex analysis have analogues in quaternionic analysis. In particular, there
are analogues of complex holomorphic functions called (left- and right-) regular functions.
Cauchy’s integral formula for complex holomorphic functions f (z)

1 f (z)
I
f (z0 ) = dz
2πi z − z0
extends to the quaternionic setting, and its analogues for (left- and right-) regular functions
are usually referred to as Cauchy-Fueter formulas. For modern introductions to quaternionic
analysis see, for example, [S, CSSS].
Complex numbers C and quaternions H are special cases of Clifford algebras. (For an
elementary introduction to Clifford algebras see, for example, [G].) There is a further ex-
tension of complex and quaternionic analysis called Clifford analysis. For Clifford algebras
associated to positive definite quadratic forms on real vector spaces, Clifford analysis is very
similar to complex and quaternionic analysis (see, for example, [BDS, DSS, GM] and ref-
erences therein). Furthermore, J. Ryan has initiated the study of Clifford analysis in the
setting of complex Clifford algebras [R1, R2].
If a Clifford algebra is associated to a quadratic form that is not positive definite, the
function that should serve as the reproducing kernel in the Cauchy type formula has singular-
ities that always intersect the contour of integration, thus rendering a potential reproducing
integral formula meaningless. For this reason, most developments of real Clifford analysis

Department of Mathematics, Indiana University, Rawles Hall, 831 East 3rd St, Bloomington, IN 47405

Undergraduate student at Cornell University, Ithaca, NY 14850

1
so far have been in the setting of Clifford algebras associated to positive definite quadratic
forms.
We develop Clifford analysis in the setting of Clifford algebras associated to indefinite
non-degenerate quadratic forms by treating the relevant Clifford algebras as real subalgebras
of complex Clifford algebras and, in a certain sense, restricting the complex Dirac operators
and functions to the appropriate real subspaces. This is done in complete parallel with
K. Imaeda’s approach in [I], where he first extends classical quaternionic analysis to the
algebra of biquaternions HC = H ⊗R C and then “restricts” to the Minkowski space M by
realizing M as a real form of HC .
Similarly, the works [L, FL] extend classical quaternionic analysis to split quaternions
HR (also known as coquaternions). The algebra HR is isomorphic to the algebra of real 2 × 2
matrices and is another example of a Clifford algebra, this time associated to an indefinite
non-degenerate quadratic form on R2 . The issue of singularities intersecting the contour
of integration is resolved in two different ways, leading to two types of integral formulas.
The first method uses biquaternions HC , which contain the split quaternions HR as a real
subalgebra. By considering the holomorphic extension of a regular function from HR into
HC and deforming the contour of integration so it no longer intersects the singularities,
one obtains the first version of Cauchy’s formulas for regular functions in HR . The second
method involves inserting a purely imaginary term iεkX − X0 k2 into the denominator of the
reproducing kernel to prevent singularities, then showing that the limit as ε → 0 exists and
produces another version of Cauchy’s formulas for regular functions in HR .
Methods for dealing with non-positive definite signature setting developed in [L, FL]
were used in [RCW] and [P]. Thus [RCW] extends the results of classical Clifford analysis to
the setting of what the authors call split-quaternionic Hermitian Clifford analysis. And [P]
extends analysis over octonions to split octonions. Both papers have to deal with exactly the
same issue of the set of singularities intersecting the contour of integration, and this issue is
resolved in the same way as in [L, FL]. In this paper we use methods of [L, FL] to extend
constructions of classical Clifford analysis to the case of indefinite non-degenerate quadratic
forms. In a larger context, our results suggest that the methods of handling singularities
developed in [L, FL] are quite general.
We begin by recalling in Section 2 the construction of the universal Clifford algebras.
These constructions are done in both the real and complex cases, and we regard the real
Clifford algebra as a subalgebra of the complex Clifford algebra. Following this, in Section
3 we define Dirac operators and introduce left- and right-monogenic functions as functions
annihilated by Dirac operators. Both real and complex cases are considered. The complex
case was mostly developed by J. Ryan [R1, R2], and the real case is, in a certain sense,
the restriction of the complex operators and functions to the real subspace. In Section 4 we
introduce Clifford-valued differential forms Dn z, Dp,q x and establish some of their properties.
These forms appear in the statements of Cauchy’s formulas for monogenic functions. We
also state the classical Cauchy’s integral formulas for left- and right-monogenic functions in
the positive definite case (Theorem 16). In Section 5 we state and prove the first version
of Cauchy’s integral formulas for left- and right-monogenic functions (Theorem 20). This
version requires having holomorphic extension of the monogenic functions to the complex
space and utilizes a deformation of the contour of integration to avoid the set of singularities.
In Section 6 we prove the second version of Cauchy’s integral formulas for left- and right-

2
monogenic functions (Theorem 23). This version involves inserting a purely imaginary term
iεkX − X0 k2 into the denominator of the reproducing kernel to prevent singularities, then
taking limit as ε → 0+ . We first establish Theorem 23 up to a constant coefficient (26).
Then we pin down the value of the coefficient using the first version of integral formulas
(Theorem 20), which is already established (Lemma 33).
This research was made possible by the Indiana University, Bloomington, Math REU
(research experiences for undergraduates) program, funded by NSF Award #1757857. We
would like to thank Prof. Chris Connell for organizing and facilitating this wonderful pro-
gram. We would also like to thank Ms. Mandie McCarthy for her administrative work, the
various professors who gave talks, and the other REU students for their company.

2 Clifford Algebras
In this section we recall basics of Clifford Algebras and establish notations. For an elementary
introduction to Clifford algebras see, for example, [G].
Let V be an n-dimensional real vector space, and Q a quadratic form on V . The form Q
uniquely extends to a bilinear form on V . Diagonalizing Q, we have that there exist integers
p, q and orthogonal basis {e1 , . . . , en } of (V, Q), such that

1
 1 ≤ j ≤ p;
Q(ej ) = −1 p + 1 ≤ j ≤ p + q;

0 p + q < j ≤ n.

By Sylvester’s Law of Inertia, we the numbers p and q are independent of basis chosen. The
ordered pair (p, q) is the signature of (V, Q). From this point on we restrict our attention to
non-degenerate quadratic forms Q, in which case p + q = n. If Q is such a form with q = 0
or p = 0, then it is positive or negative definite quadratic form respectively.
Recall the standard construction of the universal Clifford algebra associated to (V, Q) as
a quotient of the tensor algebra. We start with the tensor algebra over V ,
O
V = R ⊕ V ⊕ (V ⊗ V ) ⊕ (V ⊗ V ⊗ V ) ⊕ . . . ,
consider a set of elements of the tensor algebra
O
S = {v ⊗ v + Q(v) : v ∈ V } ⊂ V,
N
and let (S) denote the ideal of V generated by the elements of S. Then the universal
Clifford algebra associated to (V, Q) can be defined as a quotient
O
AQ = V /(S). (1)
We note the sign convention chosen for elements of the ideal is not standard in the literature,
resulting in a possible interchange of p and q. Let e0 ∈ AQ denote the multiplicative identity
of the algebra. If v1 and v2 are orthogonal elements of (V, Q),

v1 ⊗ v2 + v2 ⊗ v1 = (v1 + v2 ) ⊗ (v1 + v2 ) − v1 ⊗ v1 − v2 ⊗ v2
= −Q(v1 + v2 ) + Q(v1 ) + Q(v2 ) = 0. (2)

3
Thus AQ is a finite-dimensional algebra over R generated by e1 , . . . , en , and let Ap,q denote
the algebra AQ constructed from (V, Q), where Q has signature (p, q). We fix an orthogonal
basis
{e1 , e2 , . . . , ep , ẽp+1, . . . , ẽp+q }
of V such that Q(ej ) = 1 and Q(ẽj ) = −1 for all applicable j. In the positive definite case
the basis of V is given by {e1 , e2 , . . . , en }, and we let An = An,0. We have the following
relations for 1 ≤ i, j ≤ n with i 6= j,

e20 = e0 , e2i = −e0 , ẽ2j = e0 , e0 ei = ei e0 = ei , e0 e˜i = e˜i e0 = e˜i , (3)


ei ej = −ej ei , ei e˜j = −e˜j ei , e˜i e˜j = −e˜j e˜i . (4)

We consider subsets B ⊆ {1, . . . , n}. If B has k > 0 elements,

B = {i1 , i2 , . . . , ik } ⊆ {1, . . . , n}

with i1 < i1 < · · · < ik , define

eB = ei1 i2 ...ik = ei1 ⊗ ei2 ⊗ · · · ⊗ ẽik (5)

or, more precisely, eB is the image of this tensor product in AQ . If B is empty, we set e∅ be
the identity element e0 . These elements

eB : B ⊆ {1, 2, . . . , n}

form a vector space basis of Ap,q over R. We will be especially concerned with the space
R ⊕ V , which we identify with the vector subspace Rp+q+1 ⊂ Ap,q spanned by

{e0 , e1 , e2 , . . . , ep , ẽp+1 , . . . , ẽp+q }.

Any X ∈ Rp+q+1 can be written as


p p+q
X X
X= xj ej + x̃j ẽj .
j=0 j=p+1

n+1
Pnparticular, in the positive definite case any X ∈ R
In ⊂ An can be expressed as X =
j=0 xj ej .
We can also construct a complex universal Clifford algebra from a quadratic space over
C. Let V C be an n-dimensional complex vector space with quadratic form Q. Diagonalizing
Q, we have that there exists an orthogonal basis {e1 , . . . , en } of V C and integer p such that
Q(ej ) = 1 if 1 ≤ j ≤ p and Q(ej ) = 0 otherwise. The form Q is non-degenerate if and only
if p = n, and we only consider this case. We perform a similar construction to (1): form the
tensor algebra over V C
O
V C = C ⊕ V C ⊕ (V C ⊗ V C ) ⊕ (V C ⊗ V C ⊗ V C ) ⊕ . . . ,

consider a set O
S C = {v ⊗ v + Q(v) : v ∈ V C } ⊂ V C,

4
and let (S C ) denote the ideal generated by S C . Then the complex universal Clifford algebra
associated to (V C , Q) is a quotient
O
ACQ = V C /(S C ).

We also use notation ACn for this algebra. The analogues of (2) and (3)-(4) remain valid for
ACn , and the products eB defined as in (5) form a basis of ACn .
We can also consider ACn as a real algebra generated by

e0 , e1 , e2 , . . . , en , ie0 , ie1 , ie2 , . . . , ien .

By the universality property of Ap,q , the natural inclusion ι : V ֒→ V C defined on basis


vectors by
ι(ej ) = ej , 1 ≤ j ≤ p, ι(ẽj ) = iej , p + 1 ≤ j ≤ p + q, (6)
extends to an injective R-algebra homomorphism ι : Ap,q ֒→ ACp+q . Thus, we can consider
Ap,q as a unital real subalgebra of ACp+q .
We identify the C-span of e0 , e1 , . . . , en with Cn+1 ⊂ ACn . On Cn+1 we have the Clifford
conjugation defined by
n
X n
X
+
Z = z0 e0 + zj ej 7→ Z = z0 e0 − zj ej
j=1 j=1

and the complex conjugation defined by


n
X n
X
Z = z0 e0 + zj ej 7→ Z̄ = z¯0 e0 + z¯j ej .
j=1 j=1

(These conjugations can be extended to all of ACp+q , but for the purposes of this paper
we do not need this.) The Clifford conjugation fixes the C-span of e0 identified with C,
and the complex conjugation fixes the R-span of e0 , . . . , en identified with Rn+1 , and these
conjugations can be viewed as reflecting over the respective subspaces.
These two conjugations commute and lead to two useful quadratic forms on Cn+1 ⊂ ACn .
The first such form is n
X
+ +
N(Z) = ZZ = Z Z = zj2
j=0

(note that it is complex valued). The corresponding bilinear form is


n
1 X
hZ, W i = (Z + W + ZW + ) = zj wj . (7)
2 j=0

We denote by NnC the null cone of N(Z):

NnC = {Z ∈ Cn+1 ; N(Z) = 0},

5
then all Z ∈ Cn+1 \ NnC are invertible with inverse given by Z −1 = N(Z)−1 Z + . The other
quadratic form is real valued:
n
21 1 X
kZk = (Z Z̄ + + Z̄Z + ) = (Z̄ + Z + Z + Z̄) = |zj |2 .
2 2 j=0

As was mentioned earlier, we consider Ap,q as a real subalgebra of ACp+q . We describe the
restrictions of these two conjugations and two quadratic forms to Rp+q+1 ⊂ Ap,q ⊂ ACp+q :
p p+q p p+q
X X X X
+
X = x0 e0 + xj ej + x̃j ẽj 7→ X = x0 e0 − xj ej − x̃j ẽj ,
j=1 j=p+1 j=1 j=p+1

p p+q p p+q
X X X X
X = x0 e0 + xj ej + x̃j ẽj 7→ X̄ = x0 e0 + xj ej − x̃j ẽj , (8)
j=1 j=p+1 j=1 j=p+1

p p+q
X X
+ +
N(X) = XX = X X = x2j − x̃2j ,
j=0 j=p+1
p q
1
2 + + 1 + +
X
2
X
kXk = (X X̄ + X̄X ) = (X̄ X + X X̄) = xj + x̃2j . (9)
2 2 j=0 j=p+1

The bilinear form on Rp+q+1 corresponding to N(X) is


p p+q
1 + +
X X
hX, Y i = (X Y + XY ) = xj yj − x̃j ỹj .
2 j=0 j=p+1

In line with the complex case, we consider the null cone

Np,q = {X ∈ Rp+q+1 ; N(X) = 0},

then all X ∈ Rp+q+1 \ Np,q are invertible with inverse given by X −1 = N(X)−1 X + . Finally,
we note that in the positive definite case Nn,0 = {0}.

3 Dirac Operators, Monogenic Functions and Green’s


Functions
In this section we recall the definitions of Dirac operators, monogenic functions and Green’s
functions in the setting of complex Clifford algebras that were originally introduced in [R1].
Then, using the inclusion Rp+q+1 ⊂ Ap,q ⊂ ACp+q , we restrict these notions to define their
analogues in the setting of real Clifford algebras Ap,q . When we apply these restricted Dirac
operators to the restricted Green’s functions, we obtain special monogenic functions that
will serve as reproducing kernels of Cauchy’s integral formulas in the Ap,q setting.

6
We introduce linear differential operators on Cn+1
n n
∂ X ∂ ∂ X ∂
∇+
C = e0 + ej and ∇C = e0 − ej
∂z0 j=1 ∂zj ∂z0 j=1 ∂zj

which may be applied to functions on the left and on the right. Let C be the complex
Laplacian
n
X ∂2
C = 2
.
j=0
∂z j

Then
∇C ∇+ +
C f = ∇C ∇C f = C f and g∇C ∇+ +
C = g∇C ∇C = C g, (10)
where f is a holomorphic function on Cn+1 with values in ACn or a left ACn -module, and g is
a holomorphic function on Cn+1 with values in ACn or a right ACn -module.
Definition 1. Let U ⊆ Cn+1 be an open set, and MnC a left ACn -module. A holomorphic
function f : U → MnC is called complex left-monogenic if
n
∂f X ∂f
∇+
C f = e0 + ej =0
∂z0 j=1 ∂zj

at all points in U.
Similarly, let M̃nC be a right ACn -module, then a holomorphic function g : U → M̃nC is
called complex right-monogenic if
n
∂g X ∂g
g∇+
C = e0 + ej = 0
∂z0 j=1
∂zj

at all points in U.
The factorization (10) leads to a natural method of constructing complex left- and right-
monogenic functions. If ϕ : U → MnC is complex harmonic (i.e., C ϕ = 0), with U and
MnC as in the definition, then ∇C f is complex left-monogenic. Similarly, if ϕ̃ : U → M̃nC is
complex harmonic, ϕ̃∇C is complex right-monogenic.
We restrict these definitions to the subalgebra Ap,q ⊂ ACp+q . Thus we introduce linear
differential operators
p p+q
∂ X ∂ X ∂
∇+
p,q = e0 + ej − ẽj and
∂x0 j=1 ∂xj j=p+1 ∂ x̃j
p p+q
∂ X ∂ X ∂
∇p,q = e0 − ej + ẽj
∂x0 j=1 ∂xj j=p+1 ∂ x̃j

which may be applied to functions on the left and on the right. Let p,q be a wave operator
on Rp+q+1 ,
p p+q
X ∂2 X ∂2
p,q = − .
j=0
∂x2j j=p+1 ∂ x̃2j

7
Then
∇p,q ∇+ +
p,q f = ∇p,q ∇p,q f = p,q f and g∇p,q ∇+ +
p,q = g∇p,q ∇p,q = p,q g, (11)
where f is a C 2 function on Rp+q+1 with values in Ap,q or a left Ap,q -module, and g is a C 2
function on Rp+q+1 with values in Ap,q or a right Ap,q -module.
Definition 2. Let U ⊂ Rp+q+1 be an open set, and Mp,q a left Ap,q -module. A C 1 function
f : U → Mp,q is called (p, q)-left-monogenic if
p p+q
∂f X ∂f X ∂f
∇+
p,q f = e0 + ej − ẽj =0
∂x0 j=1 ∂xj j=p+1 ∂ x̃j

at all points in U.
Similarly, let M̃p,q be a right Ap,q -module, then a C 1 function g : U → M̃p,q is called
(p, q)-right-monogenic if
p p+q
∂g X ∂g X ∂g
g∇+
p,q = e0 + ej − ẽj = 0
∂x0 j=1
∂xj j=p+1
∂ x̃j

at all points in U.
The factorization (11) leads to a natural method of constructing (p, q)-left and right-
monogenic functions. If ϕ : U → Mp,q satisfies p,q ϕ = 0, with U and Mp,q as in the
definition, then ∇p,q f is (p, q)-left-monogenic. Similarly, if ϕ̃ : U → M̃p,q satisfies p,q ϕ̃ = 0,
ϕ̃∇p,q is (p, q)-right-monogenic.
The following case is an important example of such a construction, in which we consider
a Green’s function on Cp+q+1 ⊂ AC p+q and a suitable restriction to R
p+q+1
⊂ Ap,q .
Definition 3. For odd n ≥ 2, we define the following function Hn (Z): Cn+1 \ NnC → C by
1 1
Hn (Z) = n−1 = Pn n−1 .
N(Z) 2 ( j=0 zj2 ) 2
We note that if n is even, this may not be well defined. In this case, we introduce a
domain
Cn+1
G = Cn+1 \ {Z ∈ Cn+1 ; N(Z) ∈ R, N(Z) ≤ 0}
1
and choose a branch of N(Z) 2 : Cn+1
G → C with values in the right half-plane {w ∈ C; Re w >
1 1−n
0}. Then Hn (Z) = N(Z) 2 is a well defined function Cn+1
G → C. From now on we
n+1
restrict the domain of Hn (Z) to CG regardless of the parity of n, as the same proofs hold
in either case. Observe that on this domain Hn (Z) is complex harmonic, and, since Hn (Z) is
scalar-valued, ∇C Hn (Z) = Hn (Z)∇C is complex left- and right-monogenic. Thus, we obtain
a function on Cn+1
G with values in ACn that is complex left- and right-monogenic:
Lemma 4. An ACn -valued function Gn (Z) defined on Cn+1
G as
Pn
1 z0 − j=1 zj ej Z+
Gn (Z) = ∇C Hn (Z) = Pn n+1 = n+1
1−n ( j=0 zj2 ) 2 N(Z) 2
is complex left- and right-monogenic.

8
Restricting this function to Rp+q+1 ⊂ Ap,q as was done with Dirac operators earlier,
we obtain the following solution to the wave equation p,q ϕ = 0 and the corresponding
(p, q)-left- and (p, q)-right-monogenic functions. Let
p+q+1
RG = Rp+q+1 \ {X ∈ Rn+1 ; N(X) ≤ 0} = Rp+q+1 ∩ Cp+q+1
G .

Definition 5. For all p + q ≥ 2, define the functions Hp,q (X) : Rp+q+1


G → R and Gp,q (X) :
p+q+1
RG → Ap,q by

1 1
Hp,q (X) = Hp+q (X) = p+q−1 =  p+q−1 ,
p
N(X) 2 x2j − p+q x̃2j 2
p+q+1
P P
RG
j=0 j=p


1
Gp,q (X) = Gp+q (X) = ∇p,q Hp,q (X)
p+q+1
RG 1−p−q
x0 − pj=1 xj ej − p+q
P P
j=p+1 x̃j ẽj X+
= P  p+q+1 = p+q+1 .
p 2
Pp+q 2 N(X)
j=0 xj − j=p+1 x̃j
2 2

Consequently, p,q Hp,q (X) = 0 and Gp,q is (p, q)-left- and right-monogenic.

4 Differential Forms
In this section we introduce the Clifford-valued differential n-forms Dn z and Dp,q x that will
appear in the statement of Cauchy’s integral formulas. We also prove several properties of
these forms. At the end of the section we state the classical Cauchy’s integral formulas for
monogenic functions. The complex case originally was developed in [R1].
Definition 6. Let dVC be the n + 1 complex holomorphic form on Cn+1 :

dVC = dz0 ∧ dz1 ∧ · · · ∧ dzn ; (12)

it is normalized so that dVC (e0 , e1 , e2 , . . . , en ) = 1.


Recall the bilinear form (7) on Cn+1 .
Definition 7. Let Dn z be the unique Cn+1 -valued complex holomorphic n-form on Cn+1 such
that, for all Z0 , Z1 , . . . , Zn ∈ Cn+1 , we have


Z0 , Dn z(Z1 , Z2 , . . . , Zn ) = dVC (Z0 , Z1 , Z2 , . . . Zn ). (13)

Explicitly, we can express Dn z as a sum of n + 1 terms by substituting basis vectors into


(13) yielding
Xn
Dn z = (−1)j ej dẑj , (14)
j=0

where
dẑj = dz0 ∧ dz1 ∧ · · · ∧ dzj−1 ∧ dzj+1 ∧ . . . dzn .

9
Lemma 8. Let U ⊂ Cn+1 be an open set and let f : U → ACn , g : U → M̃nC be holomorphic
functions, where M̃nC is a right ACn -module. Then,

d(gDn zf ) = (g∇+ +
C )f dVC + g(∇C f )dVC and d(gDn z) = (g∇+
C )dVC . (15)

Similarly, if f : U → MnC and g : U → ACn are holomorphic functions, where MnC is a left
ACn -module,

d(gDn zf ) = (g∇+ +
C )f dVC + g(∇C f )dVC and d(Dn zf ) = (∇+
C f )dVC .

Corollary 9. A holomorphic function f on U is complex left-monogenic if and only if the


n-form Dn zf is closed on U. Similarly, a holomorphic function g on U is complex right-
monogenic if and only if the n-form gDn z is closed on U.

This corollary may be used to give an alternative definition of complex left- and right-
monogenic functions.
Recall that the generators of the real universal Clifford algebra Ap,q and complex universal
Clifford algebra ACp+q are related by (6). Therefore, restricting to Rp+q+1 , we have


dzj (xj ej ) = xj = dxj (xj ej ), hence dzj = dxj if 0 ≤ j ≤ p;
R p+q+1

dzj (x̃j ẽj ) = ix̃j = idx̃j (x̃j ẽj ), hence dzj = idx̃j if p + 1 ≤ j ≤ p + q.
Rp+q+1

Definition 10. We define dVp,q to be the restriction of dVC to Rp+q+1:



= iq dx0 ∧ · · · ∧ dxp ∧ dx̃p+1 ∧ · · · ∧ dx̃p+q ;

dVp,q = dVC (16)
Rp+q+1

it is normalized so that dVp,q (e0 , e1 , . . . , ep , ẽp+1 , . . . , ẽp+q ) = iq .

Definition 11. Let Dp,q x be the restriction of Dp+q z to Rp+q+1, so that we have, for all
(X0 , X1 , . . . , Xp+q ) ∈ Rp+q+1 ,


X0 , Dp,q x(X1 , X2 , . . . , Xp+q ) = dVp,q (X0 , X1 , . . . Xp+q ).

Explicitly,
p
X p+q 
X
Dp,q x = i q j
(−1) ej dx̂j − j ˆ
(−1) ẽj dx̃j , (17)
j=0 j=p+1

where

dx̂j = dx0 ∧ · · · ∧ dxj−1 ∧ dxj+1 ∧ · · · ∧ dxp ∧ dx̃p+1 ∧ · · · ∧ dx̃p+q if 0 ≤ j ≤ p,


dx̃ˆj = dx0 ∧ · · · ∧ dxp ∧ dx̃p+1 ∧ · · · ∧ dx̃j−1 ∧ dx̃j+1 ∧ · · · ∧ dx̃p+q if p + 1 ≤ j ≤ p + q.

We have the following real analogue of Lemma 8; its proof is the same as in the complex
case.

10
Lemma 12. Let U ⊂ Rp+q+1 be an open set and let f : U → Ap,q , g : U → M̃p,q be C 1
functions, where M̃p,q is a right Ap,q -module. Then,

d(gDp,q xf ) = (g∇+ +
p,q )f dVp,q + g(∇p,q f )dVp,q and d(gDp,q x) = (g∇+
p,q )dVp,q .

Similarly, if f : U → Mp,q and g : U → Ap,q are C 1 functions, where Mp,q is a left Ap,q -
module,

d(gDp,q xf ) = (g∇+ +
p,q )f dVp,q + g(∇p,q f )dVp,q and d(Dp,q xf ) = (∇+
p,q f )dVp,q .

Corollary 13. A C 1 function f on U is (p, q)-left-monogenic if and only if the n-form Dp,q xf
is closed on U. Similarly, a C 1 function g on U is (p, q)-right-monogenic if and only if the
n-form gDp,q x is closed on U.

Again, this corollary may serve as an alternative definition of (p, q)-left- and (p, q)-right-
monogenic functions.
We orient Rp+q+1 ⊂ Ap,q so that

{e0 , e1 , e2 , . . . , ep , ẽp+1 , . . . , ẽp+q }

is a positively oriented basis. For an open subset U ⊂ Rp+q+1 with piecewise smooth bound-
ary ∂U, at each smooth boundary point X ∈ ∂U, let

nX = n0 e0 + n1 e1 + · · · + np ep + ñp+1 ẽp+1 + · · · + ñp+q ẽp+q

be an outward pointing normal unit vector to U at X (with respect to (9)). Then we can
orient the boundary ∂U so that a basis {v1 , . . . , vp+q } of the tangent space TX U is positively
oriented if and only if {nX , v1 , . . . , vp+q } is a positively oriented basis of Rp+q+1 . Let dSp,q
be the contraction of nX with dVp,q , so that

dSp,q (v1 , . . . , vp+q ) = dVp,q (nX , v1 , . . . , vp+q )

whenever v1 , . . . , vp+q are vectors in the tangent space TX U of ∂U at X. Explicitly,


p
X p+q 
X
dSp,q = i q j
(−1) nj dx̂j + jˆ
(−1) nj dx̃j . (18)
j=0 j=p+1

Lemma 14. With n̄X denoting the complex conjugate of nX (recall (8)), the restriction


Dp,q x = n̄X dSp,q .
∂U

Proof. By definition, nX is orthogonal to the tangent space TX ∂U. Thus, for every vector
X = x0 e0 + x1 e1 + · · · + xp ep + x̃p+1 ẽp+1 + · · · + x̃p+q ẽp+q in TX ∂U, we have
p p+q
X X
nj xj + ñj x̃j = 0,
j=0 j=p+1

11
and hence
n0 dx0 + · · · + np dxp + ñp+1 dx̃p+1 + · · · + ñp+q dx̃p+q = 0.
Without loss of generality we may suppose n0 6= 0, then we can isolate dx0 :
1
dx0 = − (n1 dx1 + n2 dx2 + · · · + np dxp + ñp+1 dx̃p+1 + · · · + ñp+q dx̃p+q ). (19)
n0
We introduce the symbol dx̌j denoting wedge products of all but dx0 , dxj :

dx̌j = dx1 ∧ · · · ∧ dxj−1 ∧ dxj+1 ∧ · · · ∧ dxp ∧ dx̃p+1 ∧ · · · ∧ dx̃p+q if 1 ≤ j ≤ p,


dx̃ˇj = dx1 ∧ · · · ∧ dxp ∧ dx̃p+1 ∧ · · · ∧ dx̃j−1 ∧ dx̃j+1 ∧ · · · ∧ dx̃p+q if p + 1 ≤ j ≤ p + q.

Then

dxj ∧ dx̌j = (−1)j−1dx̂0 if 1 ≤ j ≤ p,


dx̃j ∧ dx̃ˇj = (−1)j−1dx̂0 if p + 1 ≤ j ≤ p + q.

Substituting (19) into the expression (18) for dSp,q and using these notations,
 p  n2  p+q  ñ2  
j j
X X
q j j
dSp,q = i n0 dx̂0 + (−1) − dxj ∧ dx̌j ) + (−1) − dx̃j ∧ dx̌j
j=1
n0 j=p+1
n 0

p p+q
iq X 2 iq
X 
2
= n + ñ dx̂0 = dx̂0 ,
n0 j=0 j j=p+1 j n0

since nX is a unit vector. We perform a similar computation for Dp,q x substituting (19) into
(17) to get
p
X p+q 
X
q j
(−1) ẽj dx̃ˆj
j

Dp,q x =i (−1) ej dx̂j −
∂U j=0 j=p+1
 p  n  p+q  ñ  
j j
X X
= iq n0 dx̂0 + j
(−1) − ej dxj ∧ dx̌j − (−1) − j
ẽj dx̃j ∧ dx̃ˇj
j=1
n0 j=p+1
n 0

p p+q
iq X iq
 X 
= nj ej − ñj ẽj dx̂0 = n̄X dx̂0 = n̄X dSp,q .
n0 j=0 j=p+1
n0

Let Kr and Sr be the boundaries of the sets {X ∈ Rp+q+1; N(X) ≤ r} and {X ∈


Rp+q+1
; kXk2 ≤ r} respectively. Note that the outward pointing normal vectors at X will
be respectively X̄/kXk and X/kXk, so we get the following analogue of Lemma 3 of [L]:
Corollary 15.

X X̄ X̄
Dp,q x = dSp,q , Dp,q x = dSp,q = dSp,q .
Kr kXk Sr kXk r

12
These two expressions are the same in the positive definite case (when q = 0).
We state the classical Cauchy’s integral formulas for left- and right-monogenic functions
in the positive definite case (q = 0). For details and proofs see, for example, [BDS, DSS, GM].
Recall that in this case we write An instead of An,0 .

Theorem 16. Let n ≥ 2. Let U ⊂ Rn+1 ⊂ An be an open bounded set with piecewise C 1
boundary ∂U, and let Mn be a left An -module. Suppose that an Mn -valued function f is
(n, 0)-left-monogenic function on a neighborhood of the closure U . We have:
(
ωn f (X0 ) if X0 ∈ U,
Z
Gn,0 (X − X0 )Dn,0 xf (X) =
∂U 0 if X0 ∈
/ U,

where ωn denotes the n-dimensional volume of the unit n-sphere in Rn+1 .


Similarly, if M̃n is a right An -module and g is an M̃n -valued function that is (n, 0)-right-
monogenic on a neighborhood of U , then
(
ωn g(X0) if X0 ∈ U,
Z
g(X)Dn,0xGn,0 (X − X0 ) =
∂U 0 if X0 ∈
/ U.

5 First Cauchy’s Integral Formula


In this section we present an integral formula for (p, q)-monogenic functions that have com-
plex holomorphic extension to some open neighborhood in Cp+q+1. The statement of the
formula involves a one-parameter deformation map hε that is used to deform the contour of
p+q+1
integration so that it does not cross the singularities of N(X − X0 )− 2 .

Definition 17. For 0 ≤ ε ≤ 1, let hε : Cp+q+1 → Cp+q+1 be a linear map defined by


p+q p q+1
X X X
Z= zj ej 7→ hε (Z) = (1 + iε)zj ej + (1 − iε)zj ej .
j=0 j=0 j=p+1

Note that the map hε depends on p and q, but we suppress this dependence in order to
avoid cumbersome notations.

Corollary 18. If X = pj=0 xj ej + p+q p+q+1


⊂ Ap,q ⊂ ACp+q , then
P P
j=p+1 x̃j e˜j ∈ R

N(hε (X)) = (1 − ε2 )N(X) + 2iεkXk2 . (20)

In particular, when ε = 1,
N(h1 (X)) = 2ikXk2 .

Definition 19. For Z0 ∈ Cp+q+1 fixed, we define

hε,Z0 : Z 7→ Z0 + hε (Z − Z0 ).

13
Theorem 20. Let p+q ≥ 2. Let U ⊂ Rp+q+1 ⊂ Ap,q be an open bounded set with piecewise C 1
C
boundary ∂U, and let Mp+q be a left ACp+q -module. Suppose that an Mp+qC
-valued function f
is (p, q)-left-monogenic on a neighborhood of the closure U. Furthermore, suppose f extends
to a complex left-monogenic function f C : W C → Mp+q C
, with W C ⊂ Cp+q+1 ⊂ ACp+q an open
subset containing U. We have:
(
ωp+q f (X0 ) if X0 ∈ U,
Z
Gp+q (Z − X0 )Dp+q zf C (Z) =
(hε,X0 )∗ (∂U ) 0 if X0 ∈
/ U,

for all ε > 0 sufficiently close to 0.


C
Similarly, if M̃p+q is a right ACp+q -module and g is an M̃p+q
C
-valued function that is (p, q)-
right-monogenic on a neighborhood of U and can be extended to a complex right-monogenic
function g C : W C → M̃p+q C
, then
(
ωp+q g(X0 ) if X0 ∈ U,
Z
C
g (Z)Dp+q zGp+q (Z − X0 ) =
(hε,X0 )∗ (∂U ) 0 if X0 ∈
/ U,

for all ε > 0 sufficiently close to 0.

Remark 21. For all ε 6= 0 sufficiently close to 0 the contour of integration (hε,X0 )∗ (∂U) lies
inside W C and the integrand is non-singular, thus the integrals are well-defined. Moreover,
we will see that the value of the integral becomes constant when the parameter ε is sufficiently
close to 0.

Remark 22. If we require ε < 0, then we need to insert a factor of (−1)q into the right
hand sides of the above formulas. Thus,
(
(−1)q ωp+q f (X0 ) if X0 ∈ U,
Z
Gp+q (Z − X0 )Dp+q zf C (Z) =
(hε,X0 )∗ (∂U ) 0 if X0 ∈
/ U,

for all ε < 0 sufficiently close to 0. And similarly for the other formula.

Proof. We consider only the left-monogenic case, the right-monogenic case is similar. By
translation, we may assume that X0 = 0. Let M = supX∈∂U kXk. We restrict W C to be the
δ-neighborhood of U in Cp+q+1 for some sufficiently small δ > 0. Consider 0 < |ε| < δ/M.
If X ∈ Rp+q+1 \ {0}, then, by (20), N(hε (X)) is not a negative real, hence hε (X) ∈ Cp+q+1G .
p+q+1 C
Thus, we have that (hε )∗ (∂U) lies inside CG ∩W . By Lemma 8, the integrand is a closed
p+q+1 C
form in the region CG ∩ W , so the integral stays unchanged for all −δ/M < ε < 0 and
0 < ε < δ/M (but the values of the integral may be different on these two intervals). If
X0 = 0 ∈ / U , we have that (hε )∗ (U ) ⊆ Cp+q+1
G ∩ W C , the integrand is a closed form, and the
integral over ∂U is zero, thus we are done.
Now, suppose X0 = 0 ∈ U. Choose an r > 0 sufficiently small such that r < δ/2
and the ball Br = {X ∈ Rp+q+1 : kXk ≤ r} is contained in U. We orient the sphere
Sr = {X ∈ Rp+q+1 : kXk = r} as the boundary of Br . By Stokes’ Theorem and (15) we

14
have:
Z Z
C
Gp+q (Z)Dp+q zf (Z) − Gp+q (Z)Dp+q zf C (Z)
(hε )∗ (∂U ) (hε )∗ Sr
Z Z
C
= d(Gp+q (Z)Dp+q zf (Z)) = 0dVC = 0.
(hε )∗ (U \Br ) (hε )∗ (U \Br )

Hence, Z Z
C
Gp+q (Z)Dp+q zf (Z) = Gp+q (Z)Dp+q zf C (Z).
(hε )∗ (∂U ) (hε )∗ Sr

By varying the parameter ε from its original value to 1, we can continuously deform (hε )∗ Sr
into (h1 )∗ (Sr ) within Cp+q+1
G ∩ W C without changing the value of the integral (here we are
using r < δ/2), and so we have:
Z Z
C
Gp+q (Z)Dp+q zf (Z) = Gp+q (Z)Dp+q zf C (Z).
(hε )∗ (∂U ) (h1 )∗ Sr

Next, we rotate the sphere (h1 )∗ Sr using the map Z 7→ e−it Z, 0 ≤ t ≤ π/4. Since
N(e−it Z) = e−2it N(Z) and ke−it Zk = kZk, as t varies continuously from 0√to π/4, the
sphere (h1 )∗ Sr gets rotated inside Cp+q+1G ∩W C into S̃r√2 – the sphere of radius r 2 contained
p+q+1
in Span{e0 , e1 , . . . , ep+q } = R ⊂ Ap+q,0. Thus, we have by Stokes’ Theorem and the
classical integral formula for the definite case (Theorem 16):
Z Z
C
Gp+q (Z)Dp+q zf (Z) = Gp+q (Z)Dp+q zf C (Z)
(hε )∗ (∂U ) S̃r √2
Z
= Gp+q,0(X)Dp+q,0xf C (X) = ωp+q f (0).
S̃r √2

6 Second Cauchy’s Integral Formula


In this section we state and prove another integral formula for (p, q)-monogenic functions.
Unlike the previous version, this one that does not change the contour of integration. Instead,
we insert a purely imaginary term iεkX−X0 k2 into the denominator of the reproducing kernel
Gp,q (X − X0 ), then take limit as ε → 0+ .

Theorem 23. Let p+q ≥ 2. Let U ⊂ Rp+q+1 be a bounded open region with smooth boundary
∂U, and let Mp,q be a left Ap,q -module. Suppose that f is an Mp,q -valued function that is
(p, q)-left-monogenic in a neighborhood of the closure U . Then, for any X0 ∈ Rp+q+1 such
that ∂U intersects the cone {X ∈ Rp+q+1; N(X − X0 ) = 0} transversally, we have:
(
ωp+q f (X0 ) if X0 ∈ U,
Z
lim+ Gp,q,ε(X − X0 )Dp,q xf (X) =
ε→0 ∂U 0 if X0 ∈
/ U,

15
where Gp,q,ε is the modified Green’s function defined by

X+
Gp,q,ε(X) =  p+q+1
N(X) + iεkXk2 2

Pp Pp+q
x0 − j=1 xj ej − j=p+1 x̃j ẽj
= P  p+q+1 .
p 2
Pp+q 2
Pp 2
Pp+q 2
j=0 xj − x̃
j=p+1 j + iε( x
j=0 j + x̃
j=p+1 j ) 2

Similarly, if M̃p,q is a right Ap,q -module and g is an M̃p,q -valued function that is (p, q)-
right-monogenic in a neighborhood of U, then, for any X0 ∈ Rp+q+1 such that ∂U intersects
the cone {X ∈ Rp+q+1 ; N(X − X0 ) = 0} transversally, we have:
(
ωp+q g(X0) if X0 ∈ U,
Z
lim+ g(X)Dp,q xGp,q,ε (X − X0 ) =
ε→0 ∂U 0 if X0 ∈
/ U.
Remark 24. As in the case of the first Cauchy’s integral formula (Theorem 20), if we let
ε → 0− , we need to insert a factor of (−1)q into the right hand sides of the above formulas.
Thus, (
(−1)q ωp+q f (X0 ) if X0 ∈ U,
Z
lim Gp,q,ε(X − X0 )Dp,q xf (X) =
ε→0− ∂U 0 if X0 ∈
/ U.
And similarly for the other formula.
Proof. We change coordinates to hybrid spherical coordinates given by
Φ(ρ, ϕ1 , . . . , ϕp , θ, ψ1 , . . . , ψq−1 ) = (x0 , x1 , . . . , xp , x̃p+1 , . . . , x̃p+q ),
x0 = ρ cos θ cos ϕ1
x1 = ρ cos θ sin ϕ1 cos ϕ2
x2 = ρ cos θ sin ϕ1 sin ϕ2 cos ϕ3
..
.
xp−1 = ρ cos θ sin ϕ1 . . . cos ϕp 0 ≤ ϕ1 , ϕ2 , . . . , ϕp−1 ≤ π,
xp = ρ cos θ sin ϕ1 . . . sin ϕp 0 ≤ ψ1 , ψ2 , . . . , ψq−2 ≤ π, (21)
x̃p+1 = ρ sin θ cos ψ1 0 ≤ ϕp , ψq−1 ≤ 2π,
x̃p+2 = ρ sin θ sin ψ1 cos ψ2 0 ≤ ρ, 0 ≤ θ ≤ π2 .
..
.
x̃p+q−1 = ρ sin θ sin ψ1 . . . cos ψq−1
x̃p+q = ρ sin θ sin ψ1 . . . sin ψq−1
These are essentially the spherical coordinates for the (x0 , . . . , xp )-subspace combined with
the spherical coordinates for the (x̃p+1 , . . . , x̃p+q )-subspace. In these coordinates, we have:
p p+q
X X
N(X) = x2j − x̃2j = ρ2 cos2 θ − ρ2 sin2 θ = ρ2 cos(2θ) and
j=0 j=p+1
p p+q
2
X X
kXk = x2j + x̃2j = ρ2 cos2 θ + ρ2 sin2 θ = ρ2 .
j=0 j=p+1

16
We thus have that the null cone Np,q is the set of X in Rp+q+1 such that θ = π4 , and we
structure our argument in the vein of [L]. We use the symmetry of the change of basis matrix
with respect to p and q − 1 to calculate the determinant by means of block matrices. Let
Sn,α be the Jacobian matrix corresponding to the transformation into the standard (n + 1)-
dimensional spherical coordinates, (ρ, α1 , . . . , αn ) → (x0 , . . . , xn ) ∈ Rn+1 ,

det(Sn,α ) = ρn sinn−1 α1 sinn−2 α2 . . . sin αn−1 .

Lemma 25. The determinant of the Jacobian matrix of the change of coordinates (21) is
given by
det(DΦ) = ρ cosp θ sinq−1 θ det(Sp,ϕ ) det(Sq−1,ψ ).
Proof. We spell out the Jacobian matrix of the change of coordinates (21) and divide it into
blocks as follows:
 ∂x0 ∂x0 ∂x0 ∂x0 ∂x0 ∂x0
. . . ∂ϕ . . . ∂ψ∂xq−1 0

∂ρ ∂ϕ1 ∂ϕ2 p ∂θ ∂ψ1
 ∂x1 ∂x1 ∂x1
. . . ∂ϕ∂x1 ∂x1 ∂x1
. . . ∂ψ∂xq−1 1 
 ∂ρ ∂ϕ1 ∂ϕ2 p ∂θ ∂ψ1 
 . .. .. .. .. .. .. .. .. 
 .. . . . . . . . . 
 ∂x ∂xp ∂xp ∂xp ∂xp ∂xp ∂xp 

 p . . . . . .
 ∂ρ ∂ϕ1 ∂ϕ2 ∂ϕp ∂θ ∂ψ1 ∂ψq−1 
 
A B
DΦ =  = .
 
  Γ ∆
 
 ∂ x̃p+1 ∂ x̃p+1 ∂ x̃p+1 ∂ x̃p+1 ∂ x̃p+1 ∂ x̃p+1 ∂ x̃p+1 
 ∂ρ ∂ϕ1 ∂ϕ2
. . . ∂ϕp ∂θ ∂ψ1
. . . ∂ψq−1 
 .. .. .. . . . .. 
 
. . . . . . .
 . . . . . . . . . 
∂ x̃p+q ∂ x̃p+q ∂ x̃p+q ∂ x̃p+q ∂ x̃p+q ∂ x̃p+q ∂ x̃p+q
∂ρ ∂ϕ1 ∂ϕ2
. . . ∂ϕp ∂θ ∂ψ1
. . . ∂ψq−1

We make the following observations about the blocks A, B, Γ and ∆. First, A is the
(p + 1) × (p + 1) matrix that is obtained from Sp,ϕ by multiplying each entry by cos θ; we
denote the first column of A by ~a. Next, B is the (p + 1) × q matrix whose first column is ~b,
described below, and all other columns are zero. Similarly, Γ is the q × (p + 1) matrix whose
first column is ~c, described below, and the rest of the columns are zero. Finally, ∆ the q × q
matrix obtained from Sq−1,ψ by multiplying the first column by ρ cos θ and multiplying the
remaining columns by sin θ; we denote the first column of ∆ by d. ~
 
ρ cos θ 0 0 ...
 0 sin θ 0 . . .
A = cos θ · Sp,ϕ , ∆ = Sq−1,ψ ·  0 ,
 
 0 sin θ . . .
..
0 0 0 .
 
cos ϕ1
 sin ϕ1 cos ϕ2 
.. ~b = − ρ sin θ · ~a,
 
~a = cos θ ·  ,
 
 .  cos θ
sin ϕ1 sin ϕ2 . . . cos ϕp 
sin ϕ1 sin ϕ2 . . . sin ϕp

17
 
cos ψ1
 sin ψ1 cos ψ2 

..
 ρ cos θ
~c = sin θ ·  , d~ = · ~c.
 
 .  sin θ
sin ψ1 sin ψ2 . . . cos ψq−1 
sin ψ1 sin ψ2 . . . sin ψq−1
Lemma 26. If A is an m × m matrix, Bis an  m × n matrix, Γ is an n × m matrix, and
A B
∆ is an invertible n × n matrix, then det = det(A − B∆−1 Γ) det(∆).
Γ ∆
Proof. This standard result on block matrices is proved by factoring into triangular matrices:

A − B∆−1 Γ B∆−1
    
A B Im,m 0m,n
= .
Γ ∆ 0n,m In,n Γ ∆

Using this lemma,


 
A B
= det A − B∆−1 Γ det(∆).

det(DΦ) = det
Γ ∆

We next calculate A − B∆−1 Γ. We first compute B −1 ∆Γ, labelling the jth row of the matrix
∆−1 by r~j (horizontal vector),
    
~r1 ~c · ~r1 0 . . .
~r2     ~c · ~r2 0 . . .  
−1 ~ ~ ~
B(∆ Γ) = B  ..  ~c 0 . . .  = b 0 . . .  ..
    ~ ~
 = (~c · ~r1 )b 0 . . . .

 .    . 
~rq ~c · ~rq 0 . . .

Observe that ~c is the first column of ∆ scaled by ρsin θ


cos θ
, and the so the scalar product of it
sin θ
−1
and the first row of ∆ is ρ cos θ , as 1 is the upper left entry of ∆∆−1 . Thus, we have that
θ ~
B∆−1 Γ is a matrix with first column ρsin
cos θ
b, and the rest of columns zero. Since ~b = − ρcos
sin θ
θ
~a,
sin θ 1 2
we have that A − B∆−1 Γ is matrix A with the first column multiplied by 1 + cos 2 θ = cos2 θ .

Using the multilinearity of the determinant, and factoring out the cos θ and sin θ from the
columns of A and ∆,

1
det(DΦ) = det A − B∆−1 Γ det(∆) =

det(A) det(∆)
cos2 θ
1
= (cosp+1 θ · ρ cos θ · sinq−1 θ) det(Sp,ϕ ) det(Sq−1,ψ )
cos2 θ
= ρ cosp θ sinq−1 θ det(Sp,ϕ ) det(Sq−1,ψ ).

We return to the proof of Theorem 23. The case X0 ∈ / U is easier, so we assume X0 ∈ U.


By translation we can also assume that X0 = 0. This lemma is proved by direct computation.

18
Lemma 27. We have:
kXk2 − X̄X +
∇+
p,q Gp,q,ε = iε(p + q + 1) p+q+3 ,
2 2
N(X) + iεkXk
kXk2 − X + X̄
Gp,q,ε∇+
p,q = iε(p + q + 1) p+q+3 .
2 2
N(X) + iεkXk

By Lemma 12,

d(Gp,q,ε · Dp,q x · f ) = (Gp,q,ε∇+ + +


p,q )f dVp,q + Gp,q,ε (∇p,q f )dVp,q = (Gp,q,ε ∇p,q )f dVp,q ,

since f is (p, q)-left-monogenic. Then by Stokes’ theorem we have


Z
Gp,q,ε(X) · Dp,q x · f (X)
∂U
iε(p + q + 1)(kXk2 − X + X̄)
Z Z
=  p+q+3 f (X)dVp,q + Gp,q,ε(X) · Dp,q x · f (X), (22)
U \Br N(X) + iεkXk2 2 Sr

where Br = {X ∈ Rp+q+1; kXk ≤ r} is a ball of sufficiently small radius r centered at the


origin and Sr is its boundary sphere.
Next, we establish (in order) analogues of Lemma 17, Lemma 18, and Lemma 20 of [L].
Recall that we have selected a branch of z 1/2 with values in the right half-plane, as discussed
after Definition 3.
Lemma 28. Fix a θ0 ∈ (0, π4 ), and let p, q be non-negative integers, then we have two
distributions which send a test function g(θ) into the limits
π π
+θ0 +θ0
g(θ)dθ g(θ)dθ
Z Z
4 4
lim+  p+q+3 and lim−  p+q+3 .
ε→0 π ε→0 π
4
−θ0 cos(2θ) + iε 2
4
−θ0 cos(2θ) + iε 2

In other words, the above two limits of integrals define continuous linear functionals on the
topological vector space of smooth functions1 on the interval [ π4 − θ0 , π4 − θ0 ].
Proof. In the case of p + q = 1 mod 2, this is a consequence of Lemma 17 of [L]. Otherwise,
we modify the proof to fit the fractional case. We have that (p + q + 3)/2 = n + 1/2 for some
non-negative integer n, and we use induction on n. For the base case, n = 0, we integrate
by parts,
π π
+θ0 +θ0
g(θ)dθ sin(2θ) g(θ)dθ
Z Z
4 4

1/2 = 1/2
π
4
−θ0 cos(2θ) + iε cos(2θ) + iε
π
4
−θ0 sin(2θ)
Z π +θ0
1/2 g(θ) π/4+θ0
 
4  21 d g(θ)
= − cos(2θ) + iε + cos(2θ) + iε dθ.
sin(2θ) π/4−θ0 π
4
−θ 0
dθ sin(2θ)
1
Note that this vector space has the standard Frechét toplogy determined by the seminorms
maxθ∈[ π4 −θ0 , π4 −θ0 ] g (k) (θ), where k = 0, 1, 2, . . . .

19
1/2
As cos(2θ) + iε is integrable for all values of ε, including ε = 0, the limits as ε → 0±
exist and depend continuously on g(θ). Now we consider the case of n > 0, in which case we
can integrate by parts,
π π
+θ0 +θ0
g(θ)dθ sin(2θ) g(θ)dθ
Z Z
4 4

n+1/2 = n+1/2
π
4
−θ0 cos(2θ) + iε cos(2θ) + iε
π
4
−θ0 sin(2θ)
1/2−n π/4+θ0 Z π +θ0
cos(2θ) + iε (2n − 1)−1
 
g(θ) 4 d g(θ)
= − n− 21 dθ sin(2θ) dθ,
2n − 1 sin(2θ)

π
π/4−θ0 4
−θ 0 cos(2θ) + iε

and the result follows by induction on n.


Lemma 29.
iε(p + q + 1)(kXk2 − X + X̄)
Z
lim  p+q+3 f (X)dVp,q = 0.
ε→0 U \Br N(X) + iεkXk2 2
Proof. We write the integral in the hybrid spherical coordinates (21) and integrate out the
variables r, ϕ1 , . . . , ϕn , ψp , . . . , ψq−1 . After we do this, we retain an integral of the form
π
g(θ)dθ
Z
2
ε  p+q+3 , (23)
0 cos(2θ) + iε 2

for some function g(θ). Due to the transversality of the boundary ∂U with respect to the
null cone Np,q = {θ = π4 }, the function g(θ) is smooth at least for θ lying in some interval
R π +θ0
[ π4 − θ0 , π4 + θ0 ] with θ0 ∈ (0, π4 ). We can apply Lemma 28 to π4−θ0 . . . , and the limit of the
4
integral on the remainder of the interval exists, so the limit of (23) as ε → 0 is zero.
We introduce a constant Cp,q (it will be evaluated later in Lemma 33):
π
cosp ϑ sinq−1 ϑ
Z
2
Cp,q = lim+  p+q+1 dϑ. (24)
ε→0 0 cos(2ϑ) + iε 2

Note that this limit exists by Lemma 28.


Lemma 30.
 
rf (X)dSp,q
Z
lim+ lim+  p+q+1 = iq ωp ωq−1 Cp,q f (0).
r→0 ε→0 Sr N(X) + iεr 2 2

Proof. By Lemma 25, we can split

dSp,q = iq r p+q cosp θ sinq−1 θdθdΩp,ϕ dΩq−1,ψ ,

where
dΩn,α = ρ−n det(Sn,α )dα1 dα2 . . . dαn

20
represents the spherical volume element on the n-dimensional sphere. The factor of iq comes
from the normalization of the Euclidean volume element on this space, (16). Note that the
factors of r in the numerator and denominator of
rf (X)dSp,q
 p+q+1
N(X) + iεr 2 2

cancel. Let Fε (θ) be the following antiderivative:


θ
cosp ϑ sinq−1 ϑ
Z
Fε (θ) =  p+q+1 dϑ.
0 cos(2ϑ) + iε 2

Integrating by parts yields


Z Z θ=π/2
r −p−q f (X)dSp,q f (X) cosp θ sinq−1 θdθdΩp,ϕ dΩq−1,ψ
Z Z
q
 p+q+1 = i  p+q+1
Sr cos(2θ) + iε 2 Sq−1 Sp θ=0 cos(2θ) + iε 2
θ=π/2
∂f
Z Z Z Z
q q

=i f Fε (θ)
dΩp,ϕ dΩq−1,ψ − i Fε (θ) dθdΩϕ,p dΩψ,q−1 .
Sq−1 Sp θ=0 ∂θ
Sq−1 Sp

By the chain rule, we have ∂f


P ∂f ∂xi
∂θ
= i ∂x i ∂θ
= r · g(X), where g(X) is a smooth function.
Thus, the second term is bounded by a constant multiple of r (by Lemma 28), and so in the
limit of r → 0+ it vanishes. In the first term, as r → 0+ , f approaches f (0), and so we can
carry it out, and integrate dΩp,ϕ and dΩq−1,ψ to be the respective volumes of unit spheres to
obtain
 Z Z θ=π/2 
lim iq

lim f Fε (θ)
dΩp,ϕ dΩq−1,ψ
r→0+ ε→0+ Sq−1 Sp θ=0
θ=π/2
= iq ωp ωq−1 Cp,q f (0).

= f (0)ωp ωq−1 lim+ Fε (θ)
ε→0 θ=0

The following integrals appear as certain cross terms.


Lemma 31. For each 0 ≤ j ≤ p and 1 ≤ k ≤ q,
 
xj x̃p+k f (X) dSp,q
Z
lim lim  p+q+1 = 0.
r→0+ ε→0+ Sr
N(X) + iεr 2 2 r

Proof. Note that, as in the proof of previous lemma, once we convert to the hybrid spherical
coordinates (21), the factors of r cancel. After conversion to these coordinates, the integrand
depends on ψk as follows. For k = 1, . . . , q−1, the integrand is of the form cos ψk ·hk , where hk
is a function does not depend on ψk . And for k = q, it is of the form sin ψq−1 ·hq . Additionally,
we see from Lemma 25 that the Jacobian, as a function of ψk , is proportional to sinq−1−k ψk

21
p+q+1
for k = 1, . . . , q − 1. By absorbing the remaining factors of xj (cos(2θ) + iε)− 2 dSp,q into
∂hk
the hk , so we still have ∂ψ k
= 0, for all ε, we can rewrite the integral as
R
q−1−k
RSr cos ψk sin
 ψk hk f (X)dθdϕ1dϕ2 . . . dϕp dψ1 dψ2 . . . dψq−1 if k = 1, . . . , q − 2,
S
cos ψq−1 hq−1 f (X)dθdϕ1dϕ2 . . . dϕp dψ1 dψ2 . . . dψq−1 if k = q − 1,
R r

Sr
sin ψq−1 hq f (X)dθdϕ1dϕ2 . . . dϕp dψ1 dψ2 . . . dψq−1 if k = q.

For 1 ≤ k ≤ q − 2, when we integrate with respect to ψk , we have


Z ψk =π
cos ψk sinq−1−k ψk hk f (X)dψk
ψk =0
ψk =π Z π Z π
sinq−k ψk sinq−k ψk ∂f sinq−k ψk ∂f
= hk f (X)
− hk dψk = − hk dψk .
q−k ψk =0 0 q−k ∂ψk 0 q−k ∂ψk
∂f
The partial derivative ∂ψ k
yields a factor of r by the chain rule, as demonstrated in the proof
of Lemma 30. Then we integrate out the remaining variables and let ε → 0+ , the limit exists
by Lemma 28. As r → 0+ , the factor of r ensures the limit is zero.
For k = q − 1 and k = q, the bounds of integration are 0 ≤ ψq−1 ≤ 2π, and the same
argument works in this case as well.
Now we can prove an analogue of Lemma 20 of [L].
Lemma 32. Let Cp,q be as in (24), then
X + Dp,q xf (X)
 Z 
lim lim  p+q+1 = iq ωp ωq−1 Cp,q f (0).
r→0+ ε→0+ Sr 2
N(X) + iεr 2

Proof. Using Corollary 15, we have


X + Dp,q xf (X) X + X̄f (X) dSp,q
Z Z
p+q+1 =  p+q+1 . (25)
Sr N(X) + iεr 2
 2
Sr N(X) + iεr 2 2 r

We calculate, if X = x0 e0 + pj=1 xj ej + p+q


P P
k=p+1 x̃k ẽk ,

p p+q
2
X X
+
X X̄ = kXk − 2 xj x̃k ẽk ej .
j=0 k=p+1

We split the integral of (25) into two parts to be analyzed separately,

X + Dp,q xf (X) kXk2 f dSp,q


Z Z
 p+q+1 =  p+q+1 r
Sr N(X) + iεr 2 2 Sr N(X) + iεr 2 2

p p+q Z
X X xj x̃k ẽk ej f (X) dSp,q
−2  p+q+1 .
j=0 k=p+1 Sr N(X) + iεr 2 2 r

Applying Lemmas 30 and 31 yields the result.

22
We have proved Theorem 23 up to a constant coefficient. Indeed, combining (22) with
Lemmas 29 and 32, we obtain
Z
lim+ Gp,q,ε(X − X0 )Dp,q xf (X) = iq ωp ωq−1 Cp,q f (X0 ) if X0 ∈ U. (26)
ε→0 ∂U

And if X0 ∈ / U the same argument shows that the limit of integrals is zero. Then Theorem
23 follows from Lemma 33 below.
Lemma 33.
ωp+q
Cp,q = (−i)q .
ωp ωq−1
Proof. We prove this by applying both integral formulas to the constant function f (X) = 1
on the sphere Sp,q = {X ∈ Rp+q+1 : kXk = 1}. When we apply the first integral formula
(Theorem 20), we obtain, for all ε > 0 sufficiently close to zero,
Z + Dp+q z
Z
p+q+1 = ωp+q .
(hε )∗ (Sp,q ) N(Z) 2

We calculate the pullback of Dp+q z and rewrite this as an integral over the sphere Sp,q . Note
that hε is a linear transformation and that
(
∗ (1 + iε)dzj if 0 ≤ j ≤ p,
(hε ) dzj =
(1 − iε)dzj if p + 1 <≤ j ≤ p + q.
Using the expansion of Dp+q z in the standard basis (14), we find:
p p+q
X X
(hε )∗ Dp+q z = (−1)j ej (1 + iε)p (1 − iε)q dẑj + (−1)j ej (1 + iε)p+1(1 − iε)q−1 dẑj
j=0 j=p+1

= Dp+q z + εD̃p+q,εz,
where the form D̃p+q,ε z depends on ε polynomially and does not depend on Z. Let D̃p,q,ε x
be the restriction of D̃p+q,ε z to Rp+q+1 . Recall the conjugation (8). If X ∈ Rp+q+1 and
Z = hε (X), then, using (20),
(hε )∗ Z + = (X + iεX̄)+ = X + + iεX̄ + ,
(hε )∗ N(Z) = (1 − ε2 )N(X) + iεkXk2 = (1 − ε2 )N(X) + iε.
We have:
Z + Dp+q z (X + + iεX̄ + )(Dp,q x + εD̃p,q,εx)
Z Z
p+q+1 =  p+q+1
(hε )∗ (Sp,q ) N(Z) 2 Sp,q (1 − ε2 )N(X) + iε 2
X + Dp,q x εX̄ + D̃p,q,ε x + iεX̄ + (Dp,q x + εD̃p,q,εx)
Z Z
=  p+q+1 +  p+q+1 .
Sp,q (1 − ε2 )N(X) + iε 2 Sp,q (1 − ε2 )N(X) + iε 2
As ε → 0+ , the first integral approaches iq ωp ωq−1 Cp,q by the already established formula
(26). And the second integral approaches zero because we can factor out ε and proceed in
the same manner as we proved Lemma 29 using the hybrid coordinates (21) and Lemma
28.

23
References
[BDS] F. Brackx, R. Delanghe, F. Sommen, Clifford Analysis, Pitman, London, 1982.

[CSSS] F. Colombo, I. Sabadini, F. Sommen, D. C. Struppa, Analysis of Dirac systems


and computational algebra, Progress in Mathematical Physics, vol. 39, Birkhäuser
Boston, 2004.

[DSS] R. Delanghe, F. Sommen, V. Souček, Clifford algebra and spinor-valued functions,


Kluwer Academic Publishers Group, Dordrecht, 1992.

[FL] I. Frenkel, M. Libine, Split quaternionic analysis and the separation of the series
for SL(2, R) and SL(2, C)/SL(2, R), Advances in Math 228 (2011), 678-763; also
arXiv:1009.2532.

[G] D. Garling, Clifford Algebras: An Introduction, London Math. Society Student Texts,
Cambridge Univ. Press, 2011.

[GM] J. Gilbert, M. Murray, Clifford Algebras and Dirac Operators in Harmonic Analysis,
Cambridge Univ. Press, Cambridge, UK, 1991.

[I] K. Imaeda, “A new formulation of classical electrodynamics”, Nuovo Cimento B (11)


32 (1976), no. 1, 138-162.

[L] M. Libine, An invitation to split quaternionic analysis in I. Sabadini, F. Sommen


(Eds.), Hypercomplex Analysis and its Applications, Birkhäuser, 2011, pp. 161-180;
also arXiv:1009.2540.

[P] B. Prather, Split Octonionic Cauchy Integral Formula, Adv. Appl. Clifford Algebras
29 (2019), issue 5.

[RCW] G. Ren, L. Chen, H. Wang, Split-quaternionic Hermitian Clifford analysis, Complex


Variables and Elliptic Equations 60 (2015), 333-353.

[R1] J. Ryan, Complexified Clifford analysis, Complex Variables, Theory and Appl. 1
(1982), 119-149.

[R2] J. Ryan, Cells of harmonicity and generalized Cauchy integral formulae, Proc. Lon-
don Math. Soc. (3) 60 (1990), no. 2, 295-318.

[S] A. Sudbery, Quaternionic analysis, Math. Proc. Camb. Phil. Soc. 85 (1979), 199-225.

24

View publication stats

You might also like