You are on page 1of 18

Home Search Collections Journals About Contact us My IOPscience

Characteristics of the ultrahyperbolic differential equation governing pole density functions

This content has been downloaded from IOPscience. Please scroll down to see the full text.

1999 Inverse Problems 15 1603

(http://iopscience.iop.org/0266-5611/15/6/312)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 131.111.164.128
This content was downloaded on 05/09/2015 at 03:41

Please note that terms and conditions apply.


Inverse Problems 15 (1999) 1603–1619. Printed in the UK PII: S0266-5611(99)98935-9

Characteristics of the ultrahyperbolic differential equation


governing pole density functions

Dmitry I Nikolayev† and Helmut Schaeben‡


† Laboratory for Neutron Physics, Joint Institute for Nuclear Research, Dubna, 141 980 Moscow
Region, Russia
‡ Mathematical Geology and Computer Sciences in Geology, Freiberg University of Technology
and Mining, D-09596 Freiberg, Germany

Received 4 November 1998, in final form 19 July 1999

Abstract. In this paper we define a hyperspherical x-ray transform, characterize the range of this
transform in terms of an ultrahyperbolic differential equation which is derived here without referring
to spherical harmonics, and provide its general solution in terms of both its characteristics or
spherical harmonics. These results will aid in the solution of actual problems of texture goniometry,
i.e., the analysis of preferred crystallographic orientation in polycrystalline materials, particularly
in the solution of the inverse tomographic problem of texture goniometry.

1. Introduction

In crystallography, a probability density function f defined on the group SO(3) of rotations


is referred to as the orientation density function by volume. The corresponding pole density
function P̃h of a crystal form h is defined as the projection of the orientation density function
provided by an integral operator Ph : C 2 (SO(3)) 7→ C 2 (S 2 × S 2 ). In the past, the
majority of interest was focused on the corresponding ambiguous inverse problem to find an
orientation density function given some pole density functions which can be easily measured
in normal diffraction experiments. It almost completely slipped the attention of researchers
that pole density functions are governed by a differential equation of ultrahyperbolic type.
Exploiting this differential equation may provide the means for: (i) checking the compatibility
of experimental pole density functions; (ii) completing pole density functions incompletely
measured on some subset D ⊂ S 2 only; and (iii) directly calculating additional pole density
functions, without previous determination of a reasonable solution of the inverse problem.
Such means have been long searched for.

2. Definition of a hyperspherical x-ray transform

2.1. The projection operator

Definition. Let f ∈ C 2 (SO(3)) be a twice differentiable function defined on the group of proper
rotations. Let h, r ∈ S 2 ⊂ R3 , then the integral operator Ph : C 2 (SO(3)) 7→ C 2 (S 2 × S 2 ) is
defined as
Z
1
(Ph f )(r ) = f (g) dv(g) = P (h, r ) (1)
2π {g∈SO(3)|h=gr}
0266-5611/99/061603+17$30.00 © 1999 IOP Publishing Ltd 1603
1604 D I Nikolayev and H Schaeben

where the function P (h, r ) for a given h ∈ S 2 may be referred to as the hyperspherical x-ray
transform of f with respect to h. The geometry of the integration path {g ∈ SO(3)|h = g r }
in equation (1) will be discussed in section 2.2 and in more detail in the appendix.
Obviously,
P (−h, r ) = P (h, −r ). (2)
However, it is emphasized that for a given h ∈ S 2 , P (h, r ) 6= P (−h, r ) holds.
Furthermore, for any fixed g0 ∈ SO(3)
Ph [f (gg0−1 )](r ) = Ph [f (g)](g0 r ) (3)
Pg0 h [f (g)](r ) = Ph [f (g0 gg0−1 )](g0−1 r ) (4)
which reduces for rotationally invariant functions f (g) = f (ω(g)) which depend only on the
angle ω = arccos( 21 [tr M(g) − 1]) of the rotation g represented by the 3 × 3 matrix M(g), to
P (g0 h, r ) = P (h, g0−1 r ). (5)
Thus, if f is rotationally invariant with respect to g0 , i.e. if
f (g; g0 ) = f (ω(g, g0 )) = f (ω(gg0−1 ))
then the transform P (h, r ) defined on S 2 × S 2 is rotationally invariant with respect to
r0 = g0−1 h. It reduces to a real function P (h· g0 r ) depending on the angle < (h, g0 r ) ∈ [0, π ]
and may be thought of as being defined on [−1, 1] by virtue of the inner (dot) product h · g0 r .
It should be noted that the rotational invariance of f with respect to a given g0 is necessary
and sufficient for the rotational invariance of the h x-ray transform P (h, r ) with respect to
r0 = g0−1 h for all h ∈ S 2 ; rewriting equation (5) obviously yields
Pg0 h [f (g)](r ) = Ph [f (g)](g0−1 r ) = Ph [f (g0 gg0−1 )](g0−1 r ). (6)
Since the x-ray transformation is invertible (see below),
f (g) = f (g0 gg0−1 ) (7)
implies that f is a function of the angle ω(g) of rotation only,
f (g) = f (ω(g)) (8)
and is thus rotationally invariant. Thus, rotational invariance is preserved by the spherical
x-ray transform.
Next, for reasons originating in crystallography and Friedel’s law, which states that the
diffracting x-ray cannot distinguish between the top and bottom of the lattice planes, i.e.,
the directions h and −h cannot be distinguished in the diffraction experiment even if the
crystal itself is not centrosymmetric (cf [7], p 106), the basic crystallographic x-ray transform
P̃h : C 2 (SO(3)) 7→ C 2 (S+2 × S+2 ) is defined as
Z
1 1
(P̃h f )(r ) = f (g) dv(g) = (P (h, r ) + P (−h, r )) = P̃ (h, r ) (9)
2π {g∈SO(3)|±h=gr} 2
where S+2 denotes the upper unit hemisphere in R3 , or equivalently the projective plane H 2 .
Now, obviously,
P̃ (h, r ) = P̃ (−h, r ) = P̃ (h, −r ) (10)
that is, P̃ is an even function in both arguments h, r ∈ S 2 , and therefore the operator P̃h
essentially maps onto C 2 (S+2 × S+2 ).
In texture analysis the notation of passive rotations is preferred by most authors and in the
standard references (cf [6, 7, 21, 22]). Let KA , KB be two different coordinate systems fixed to
the sample and the crystal, respectively. The map M(g) ∈ SO(3) with M r = h corresponds
to a transformation of coordinates of a fixed unit vector, which is given by r with respect to
KA and by h with respect to KB , according to a proper rotation g : KA 7→ KB represented by
the 3 × 3 matrix M(g).
Characteristics of the spherical ultrahyperbolic differential equation 1605

2.2. Path of integration


Euler angles provide a parametrization of rotations by a triplet of three angles (α, β, γ ) of
rotations about given axes by the convention that the first by α ∈ [0, 2π ) is about the Z-axis,
the second by β ∈ [0, π] about the (new) X0 -axis, and the third by γ ∈ [0, 2π ) about the
(new) Z 00 -axis of the crystal coordinate system KB (cf [7], p 4). It should be noted that Euler
angles may be defined differently, in particular the second rotation may be defined as rotation
about the (new) Y 0 -axis (e.g. [1], p 67). Interpreting α, β as spherical coordinates of some unit
vector h the notation {h, γ } for the rotation g(α, β, γ ) was introduced in [21].
The set {g ∈ SO(3) | h = g r } may be represented as the set of two successive (passive)
rotations in terms of Euler angles. The first is a fixed rotation g1 = g1 (ϕ, ϑ, 0) given in terms
of the spherical coordinates (ϕ, ϑ) of r mapping r onto e3 = (0, 0, 1), and the second is a
variable rotation g2−1 = g2−1 (α, β, ω) given in terms of the spherical coordinates (α, β) of h
and a variable angle ω ∈ [0, 2π) mapping e3 onto h for all ω ∈ [0, 2π )
{g ∈ SO(3)|h = g r } = {g2−1 (α, β, ω)g1 (ϕ, ϑ, 0)|ω ∈ [0, 2π )} (11)
= {{h, ω}−1 {r , 0}|ω ∈ [0, 2π )} (12)
with matrix representations
!
cos ϕ cos ϑ sin ϕ cos ϑ − sin ϑ
M({r , 0}) = M(g1 ) = − sin ϕ cos ϕ 0 (13)
cos ϕ sin ϑ sin ϕ sin ϑ cos ϑ
and
M({h, ω}−1 ) = M(g2−1 (α, β, ω)) = M t (g2 (α, β, ω))
!
cos α cos β cos ω − sin α sin ω − cos α cos β sin ω − sin α cos ω cos α sin β
= sin α cos β cos ω + cos α sin ω − sin α cos β sin ω + cos α cos ω sin α sin β
− sin β cos ω sin β sin ω cos β
! !
cos α cos β − sin α cos α sin β cos ω − sin ω 0
= sin α cos β cos α sin α sin β sin ω cos ω 0
− sin β 0 cos β 0 0 1
−1 −1
= M(g22 (α, β, 0))M(g21 (0, 0, ω)). (14)
Thus, switching notation to spherical coordinates
Z
1
P (α, β; ϕ, ϑ) = f (g2−1 (α, β, ω)g1 (ϕ, ϑ, 0)) dω
2π (0,2π )
Z
1
= f (g2−1 (α, β, 0)g1 (ϕ, ϑ, ω)) dω (15)
2π (0,2π )
where the last equality follows from the decomposition stated above. From (15) it is obvious
that f ∈ C 2 (SO(3)) implies P (h, r ) ∈ C 2 (S 2 × S 2 ).
For another geometrically instructive representation of the path of integration as a great
circle restricted to S+3 ⊂ R4 , or equivalently as a line in the projective space H 3 , the reader is
referred to the appendix.

2.3. Crystallographic pole density functions


Crystallography distinguishes 32 different symmetry classes GB to which a crystal may belong,
in virtue of the occurrence of different elements of symmetry passing through a single point;
in terms of algebra, each crystal symmetry class corresponds to a finite point symmetry group
(cf [8]). Let GB = GB ∩ SO(3) denote the finite point subgroup of proper rotations associated
1606 D I Nikolayev and H Schaeben

with the crystal symmetry class GB , and #GB the total number of its elements. Due to Friedel’s
law the effective crystal symmetry is described by the point group G̃B = GB ⊗ {1, −1}, where
1 denotes identity and −1 the symmetry operation of inversion, which is also referred to as
the Laue class. Corresponding to the 32 crystal classes there exist 11 Laue classes G̃B , which
always contain the operation of inversion as an element of symmetry.
The 32 different symmetry classes GB are given by 11 purely rotational groups and 11
Laue groups; for both G̃B = GB ⊗ {1, −1} holds. Therefore, the total number of their elements
is #G̃B = 2#GB .
For the remaining ten groups
G̃B = G̃B ∩ SO(3) = (GB ⊗ {1, −1}) ∩ SO(3) 6= GB .
Considering only GB instead of G̃B means an essential loss of information.
Let g be an arbitrary element of SO(3). In the case of crystallographic symmetry the (right)
coset of orientations GB g = {gBj g|gBj ∈ GB , j = 1, . . . , #GB }, consists of orientations
which are physically indistinguishable with respect to the coordinate system KA fixed to the
sample; g is referred to as representative of the coset GB g. The cosets with respect to GB
provide a partition of SO(3). Therefore, the property
f (gBj g) = f (g), gBj ∈ GB , j = 1, . . . , #GB (16)
is imposed on f , that is, f is constant on each coset. Eventually, it will be sufficient to consider
f on the elementary set G of representatives of all cosets.
For a function f with the property (16) their corresponding projections satisfy
P (gBj h, r ) = P (h, r ). (17)
The central object of texture analysis is the orientation density function f which is used to
represent the probability by volume (not by total number) that crystals possess an orientation
g within dv(g) ⊂ G as f (g) dv(g). Its x-ray transforms according to equation (9) are referred
to as pole figures P̃h of the crystal form h in texture analysis. Pole figures are spherical density
functions used to represent the probability (by volume) that crystals possess a crystallographic
direction h within the spherical surface element ds(r ) ⊂ S+2 as P̃h (r ) ds(r ). In the context of
diffraction experiments the crystallographic direction h is usually associated with the normal
unit vector of some lattice planes with Miller indices (hkl) ∈ Z3 which denote the inverse
intercepts of the coordinate axes of KB converted to integers. The crystallographic multiplicity
of a crystal form h is the total number of symmetrically equivalent lattice planes, i.e. of
crystallographic directions hm which are symmetrically equivalent.
In a typical diffraction experiment the wavelength and the diffraction geometry of x-ray
source and counter, i.e. the Bragg angle between the incident and the diffracted beams, is chosen
and fixed. By virtue of Bragg’s equation relating wavelength, Bragg angle and interplanar
lattice spacing which actually diffracts at the Bragg angle, a lattice plane is chosen which
may be represented by its Miller indices or its normal and their symmetrically equivalents.
Rotation of the lattice plane around h changes the crystallographic orientation of the lattice
plane but leaves h fixed. Thus, the proportion, by volume, of crystal grains with h or any
of its symmetrically equivalents coinciding with a given direction r is the (continuous) sum
of proportions by volume of crystals with an orientation g such that h and r coincide, i.e.
g r = h. In a texture goniometer, rotation is actually not about h, but the sample is rotated
such that for a finite set of sampling locations rl the diffracted intensity is recorded.
These ideas will be further formalized as follows.
For an orientation density function f ∈ C 2 (SO(3)), its corresponding crystallographic
pole density function P̃h of the crystal form h = {hm |m = 1, . . . , Mh } ⊂ S 2 of multiplicity Mh
Characteristics of the spherical ultrahyperbolic differential equation 1607

corresponding to crystal-symmetrically equivalent lattice planes {(hkl)m |m = 1, . . . , Mh } ⊂


Z3 is defined for r ∈ S 2 as
1 XMh
P̃h (r ) = (P̃h f )(r ) = P̃ (hm , r ) (18)
Mh m=1
such that
Z
P̃h (r ) ds(r ) = 1
S+2

when
Z
f (g) dv(g) = 1.
SO(3)

Since for normal scattering diffraction experiments Friedel’s law applies, the set of
directions {hm , −hm |m = 1, . . . , Mh } can be identified with the set {g̃Bj h|j = 1, . . . , #G̃B } =
{gBj h, −gBj h|j = 1, . . . , #GB } and equation (18) can be rewritten as
Z
1 X 1
(P̃h f )(r ) = f (g) dv(g). (19)
#G̃B 2π {g∈SO(3)|g̃Bj h=gr}
g̃Bj ∈G̃B

Summation in (19) is thus over all directions gBj h, −gBj h, j = 1, . . . , NB , which are
symmetrically equivalent with respect to the associated Laue group G̃B . Taking (16) and
(17) into account yields
1 X
(P̃h f )(r ) = P̃ (gBj h, r ) = P̃ (h, r ). (20)
#GB gB ∈GB
j

Thus, while h denotes the set of symmetrically equivalent crystallographic directions


hm related to each other by the associated Laue group, a crystallographic pole density
function P̃h (r ) is initially defined in (18) as a superposition of corresponding projections
P̃hm , m = 1, . . . , Mh , but can eventually be thought of as a function P̃ (h, r ) of two spherical
variables h, r ∈ S 2 ⊂ R3 , where h denotes an arbitrary element of h. Thus, P̃h (r ) and P̃ (h, r )
will no longer be distinguished.
For crystallographic pole density functions it always holds that
P̃ (h, r ) = P̃ (−h, r ) = P̃ (h, −r ) (21)
independently of the class of crystal symmetry, i.e. P̃ is an even function both in h ∈ S 2 as
well as in r ∈ S 2 , while an orientation density function is generally neither even nor odd.
Summarily, without considering the restrictions imposed on hm by crystallography a pole
density function can actually be thought of as function P̃ (h, r ) : S+2 × S+2 7→ R+ because of
the antipodal symmetries.
According to equations (9) and (18), respectively, its values may be read as (the sum of)
means along some one-dimensional lines G(hm , r ) = {g ∈ SO(3) | ± hm = g r }, m =
1, . . . , Mh , of integration which in turn define the actual tomographic projection from SO(3)
onto S+2 × S+2 , or equivalently from H 3 onto H 2 × H 2 .
Pole density functions of corresponding given crystal forms h are experimentally
accessible and can be discretely sampled by diffraction (x-ray, neutron) with a texture
goniometer. The problem to determine a reasonable orientation density function from given
pole density functions gives rise to the ambiguous and ill-posed tomographic inverse problem
of texture goniometry.
1608 D I Nikolayev and H Schaeben

3. The ultrahyperbolic differential equation of texture goniometry

The conventional notation P̃h (r ) of texture analysis was dropped in favour of P̃ (h, r )
to emphasize its features of a general axis probability density function (‘allgemeine
Achsenverteilungsfunktion’ Ã(h, r ), [7], p 53) as follows.
Let 1 denote the Laplace–Beltrami operator (explicitly defined in equation (40)), and
assume f ∈ C 2 (SO(3)). Then
Z
1
1h P (h, r ) = 1h f (g2−1 (α, β, ω)g1 (ϕ, ϑ, 0)) dω
2π (0,2π )
Z
1
= 1r f (g1−1 (α, β, 0)g2 (ϕ, ϑ, ω)) dω
2π (0,2π )
= 1r P (h, r ) (22)
where the equality is obtained by simultaneously substituting h, r , g by r , h, g −1 , and using
(15).
Thus, the hyperspherical x-ray transform and hence crystallographic pole density functions
satisfy the differential equation
(1h − 1r )P (h, r ) = 0 (23)
which is referred to as an ultrahyperbolic differential equation in mathematical physics (cf [10]).
Considering the differential equation for spherical harmonics involving the Laplace–
Beltrami operator, and the representation of pole density functions in terms of spherical
harmonics given below, it was first noticed years ago by [33], and much later communicated
in [34,35], and co-workers [17,36] that crystallographic pole density functions P̃ (h, r ) satisfy
the differential equation
1r P̃ (h, r ) = 1h P̃ (h, r ). (24)
The differential equation is of ultrahyperbolic type [10, 18]. It reduces to the spherical
potential equation if the left-hand side of (23) or (24), respectively, does not depend on r , e.g.
if the orientation distribution and hence the pole distributions are uniform. It reduces to the
hyperbolic wave equation if the left-hand sides of (23) or (24), respectively, depends only on
one variable, e.g. ϑ ∈ [0, π/2] for rotationally invariant pole density functions.
Referring to the conventional notation of texture analysis where P̃h (r ) denotes the ‘direct’
pole density function of the crystal form h, and R̃r (h) the ‘inverse’ pole density function of
the sample direction r ([7], pp 76–7),
R̃r (h) = P̃ (h, r ) (25)
equation (24) may be rewritten as
1r P̃h (r ) = 1h R̃r (h).
Using an angular probability density function (‘Winkelverteilungsfunktion’ W̃hi rj (2), [7],
pp 73–6) the integral relation
Z Z
P̃hi (2, ψ) dψ = R̃rj (2, ψ 0 ) dψ 0
[r j ] [hi ]

where the notation indicates integration along small circles with centres rj and hi , respectively,
and angular distance 2, was deduced ([7], p 77) and interpreted to result for 2 = 0 in the
equation ([7], p 77)
P̃hi (rj ) = R̃rj (hi ) = Ã(hi , rj )
Characteristics of the spherical ultrahyperbolic differential equation 1609

which seems rather trivial now and which was not pursued much further, cf [15].
In the past it obviously slipped the attention of researchers that the integral relation of a
‘direct’ and ‘inverse’ pole figure is actually a special case of a spherical analogue of the mean
value theorem due to [2] (cf [10], vol 2, p 744).
The spherical Ásgeirsson’s theorem may be stated as follows.
Theorem 3.1. Let µ(h, r ; ρ), ν(h, r ; ρ), and w(h, r ; ρ, τ ) be defined as
Z
µ(h, r ; ρ) = P̃ (h0 , r ) dt (h0 ) (26)
h·h0 =ρ
Z
ν(h, r ; ρ) = P̃ (h, r 0 ) dt (r 0 ) (27)
r·r 0 =ρ
Z Z
w(h, r ; ρ, τ ) = P̃ (h0 , r 0 ) dt (h0 ) dt (r 0 ) (28)
r·r 0 =τ h·h0 =ρ

where dt denotes the one-dimensional boundary element of the small circle S 2 ∩ Sh,√2(1−ρ) =
Ch,ρ = {h0 ∈ S 2 | h0 · h = ρ} with centre h ∈ S+2 and radius sin(arccos ρ) normalized in such
a way that the integral of the function 1 is 1. Then
µ(h, r ; ρ) = w(h, r ; ρ, 1) (29)
ν(h, r ; ρ) = w(h, r ; 1, ρ) (30)
and
µ(h, r , ρ) = ν(h, r , ρ). (31)
More generally
w(h, r ; ρ, τ ) = w(h, r ; τ, ρ) (32)
the double mean value is symmetric in the radii ρ and τ .
For a proof in terms of Lie groups the reader is referred to [16], p 319, where Ásgeirsson’s
mean value theorem is proven for any function satisfying the ultrahyperbolic differential
equation on S n × S n considering the compact sphere as a special case of a rank-one symmetric
space. It should be noted that the special case presented here can be elementarily proven using
spherical harmonics.
Since equation (31) holds for all ρ ∈ (−1, 1) it provides at once a mean value theorem for
the spherical surface element h,ρ = {h0 ∈ S 2 | ρ 6 h0 · h} with centre h ∈ S+2 by integration
with respect to ρ
Z Z
µ(h, r , ρ) dρ = ν(h, r , ρ) dρ (33)
[1,δ] [1,δ]
or more explicitly
Z Z Z Z
0 0
P̃ (h , r ) dt (h ) dρ = P̃ (h, r 0 ) dt (r 0 ) dρ. (34)
[1,δ] h·h0 =ρ [1,δ] r·r 0 =ρ

4. Solving the ultrahyperbolic differential equation of texture goniometry

4.1. General solution


In complete analogy to (i) Fourier’s harmonic method and (ii) d’Alembert’s method of
characteristics to solve an ordinary hyperbolic differential equation, e.g. to solve the Cauchy
problem of mathematical physics (cf [9]), we shall derive the general solution of the
ultrahyperbolic differential equation for hyperspherical x-ray transforms both in terms of its
characteristics as well as spherical harmonics.
1610 D I Nikolayev and H Schaeben

4.2. The general solution in terms of spherical harmonics


The spherical harmonic functions Ylm (r ), l = 0, 1, . . . , m = −l, . . . , l
Ylm (r ) = Ylm (ϑ, ϕ) = exp(imϕ)Plm (cos ϑ) (35)
with the normalized associated Legendre functions of the mth order
 m
d
Pl (z) = (1 − z )
m 2 m/2
Pl (z) (36)
dzm
where Pl (z) is the lth Legendre polynomial, form an orthonormal total system of functions in
the set C 2 (S 2 ) of square integrable functions defined on S 2 .
Of special interest are the properties
Ylm (−r ) = (−1)l Ylm (r ) (37)
(Ylm )∗ (r ) = (−1)l Yl−m (r ) (38)
1Ylm (ϑ, ϕ) = −l(l + 1) Ylm (ϑ, ϕ) m = −l, . . . , l, l = 0, 1, . . . (39)
with the Laplace–Beltrami operator
1 1
1u = 2
uϕϕ +(uϑ sin ϑ)ϑ (40)
sin ϑ sin ϑ
where latitude ϑ ∈ [0, π] and longitude ϕ ∈ [0, 2π] are the usual spherical coordinates of
r ∈ S 2 . The normalized spherical harmonics are the periodic eigenfunctions of the Laplace–
Beltrami operator 1m (cf [10], ch V, VII; [32]; [14], p 41; [24], p 39; [41], p 7).

Theorem 4.1. The general solution of (1h − 1r )P (h, r ) = 0 for P ∈ C ∞ (S 2 × S 2 ) is


X
∞ X
l X
l
0 0
P (h, r ) = Clmm (Ylm )∗ (h)Ylm (r ). (41)
l=0 m=−l m0 =−l

Proof. Any function u ∈ C ∞ (S 2 × S 2 ) can be represented by its tensor product expansion into
spherical harmonics
X
∞ X
l X
∞ X
l0
0
u(h, r ) = λlm,l 0 m0 (Ylm )∗ (h)Ylm0 (r ) (42)
l=0 m=−l l 0 =0 m0 =−l 0

which simplifies for functions u(h, r ) satisfying equation (23) to


X
∞ X
l X
l
0 0
u(h, r ) = λmm
l (Ylm )∗ (h)Ylm (r ) (43)
l=0 m=−l m0 =−l

which is the form of (41) as the other terms drop out because the eigenvalues of 1 are different
for different l. Here, summation and differentiation may be exchanged due to [38]
0
λmm
l = O(l −k ) for any k ∈ N. (44)


Physical requirements like non-negativity, evenness, crystallographic or sample


symmetries impose additional constraints on the system (41) and lead to special solutions.

Theorem 4.2. For any function u ∈ C ∞ (S 2 ×S 2 ) satisfying the differential equation (23) there
exists a unique function f ∈ C ∞ (SO(3)) such that Ph f = u.
Characteristics of the spherical ultrahyperbolic differential equation 1611

Proof. The function f ∈ C ∞ (SO(3)) given by


X
∞ X
l X
l
0 0
f (g) = λmm
l Dlmm (g) (45)
l=0 m=−l m0 =−l
0
with generalized spherical harmonics Dlmm (g) defined in terms of Euler angles (α, β, γ ) as
Dlmn (g) = Dlmn (α, β, γ ) = exp(imα + inγ )Plmn (cos β) (46)
with the Jacobi polynomials
(−1)l−m in−m ((l − m)!(l + n)!)1/2
(1 − z)− 2 (1 + z)− 2
l−m l+m
Plmn (z) =
2l (l − m)!((l + m)!(l − n)!)1/2
 l−m 
d
× ((1 − z)l−m (1 + z)l+m ) (47)
dzl−m
possess the x-ray transforms (43), cf [14, 40]. The required convergence follows from
equation (44) and
 
n+q −1
max |Pl (x)| = max Pl (±1) =
mn mn
, m, n > −1, max(m, n) > − 21
n
(48)
 
dk mn 1
P (x) = O(nq ), q = max 2k + m, 2k + n, k − (49)
dx k l 2
([11], p 206). 
A more refined analysis of the smoothing properties of the x-ray transform is possible
in terms of Sobolev spaces or Fourier integral operators and will be given elsewhere. The
emphasis here is on the mere existence of a range theorem as exemplified for C ∞ .
The same result that the operator of equation (1) is invertible follows from generalizations
of Funk’s [13] original work that an even function f ∈ C(S 3 ) is uniquely determined by the
values of the integrals over all great-spheres (cf [25, 37]).
Thus, the coefficients of the general solution of the ultrahyperbolic equation for spherical
x-ray transforms in terms of harmonics are the harmonic coefficients of the function being
transformed, i.e. the solution of the ultrahyperbolic equation provides the solution of the
inverse x-ray transform problem. This possibility to solve a differential equation instead of an
inverse Radon-type problem has already been shown in [4], p 174, for a bounded domain in
two dimensions.
With respect to texture goniometry, approaching solutions of the differential equation
governing pole density functions in this generalized way of Fourier’s method is in complete
analogy to the harmonic method developed by [5, 6, 31] to solve the inverse tomographic
problem starting from the integral equation (1). However, this duality may be exploited
to improve on the conventional harmonic method of solution and actual solution, e.g., by
removing inconsistencies in the intensity data, by completing experimental pole density
functions sampled on a subset of S+2 only, i.e. generally refining the data (cf [29]).
Since a pole density function P̃ (h, r ) of any crystal form h is an even function by definition
equation (18), and the harmonics Ylm are either even or odd depending on l, it can be represented
by a sum of the even harmonics only, and its Fourier coefficients with respect to the odd
elements of the set {Ylm } vanish. Therefore, the knowledge of the general solution of the
differential equation for even functions P̃ ∈ C 2 (S 2 × S 2 ) is equivalent to the knowledge of
the even part f˜ of corresponding functions f ∈ C 2 (SO(3)) being mapped by P̃h on P̃ . If it
is possible to determine the even part f˜ of an orientation density function conforming with
1612 D I Nikolayev and H Schaeben

given crystallographic pole density functions of crystal forms h1 , . . . , hq , then it is possible to


determine the pole density function of any other crystal form, and vice versa.
The backtransform of equation (1) reads (cf [20, 26])
Z Z  Z 
θ ∂
f (g) = P (h, −g h) dh + 2 cos ν(h, g h, cos θ ) dh dθ (50)
S2 (0,π) 2 ∂ cos θ S 2
with
Z
ν(h, r , ρ) = P (h, r 0 ) dt (r 0 ) (51)
rr 0 =ρ

as given in equation (26).

4.3. The general solution in terms of characteristics


Any function u(h, r ) ∈ C 2 (S 2 × S 2 ) which depends on h · r = cos η only satisfies the
differential equation (23). Furthermore, an appropriate linear transformation of h or r ,
respectively, with constant coefficients should not change this situation. Consequently,

Theorem 4.3. The characteristics of the ultrahyperbolic differential equation are the fibres
(h · g r ) = const, where g is an arbitrary proper rotation.

Proof. Let g ∈ SO(3) and let u ∈ C 2 (S 2 × S 2 ) be of the form


u(h, r ) = u(h · g r )
The goal is to show 1h u = 1r u. First, since the Laplacian is rotation invariant,
1r [u(h · g r )] = 1r [u(h · r )]|gr , it is sufficient to prove 1h u = 1r u for u(h, r ) = u(h · r ).
Next, we use the fact [38] that, to find 1h F (h), we can use the Laplacian L on R3 evaluated
on F (x/|x|) and restrict to x = h ∈ S 2 , i.e.
1h F (h) = LF (x/|x|)|x=h .
Eventually, we apply the chain rule in R3 with
X3
∂2 h1 r1 + h2 r2 + h3 r3 h1 r1 + h2 r2 + h3 r3
2
(q q ) = −2 q q
i=1 ∂hi r 2 + r 2 + r 2 h2 + h2 + h2 (r 2 + r 2 + r 2 )( (h2 + h2 + h2 ))3
1 2 3 1 2 3 1 2 3 1 2 3

and
 
X3
∂2  h1 r1 + h2 r2 + h3 r3 h1 r1 + h2 r2 + h3 r3
q q  = −2 q q
2
i=1 ∂ri r12 + r22 + r32 h21 + h22 + h23 ( (r12 + r22 + r32 ))3 (h21 + h22 + h23 )

to find Lu(h/|h| · r /|r |), and then collect terms in u0 and u00 , and finally use h21 + h22 + h23 =
r12 + r22 + r32 = 1, which completes the proof. Again, as with the spherical Ásgeirsson theorem,
the theorem could be proven quite elementarily using spherical harmonics. 
For an explicit geometrically instructive representation of the characteristics, i.e. the set
{g ∈ SO(3)|(h · g r ) = const} we refer to the appendix.
Thus, the general solution of the ultrahyperbolic differential equation can be represented
as a sum
XX
u(h, r ) = ul (xgk ) (52)
l k
Characteristics of the spherical ultrahyperbolic differential equation 1613

with
xgk = (h · gk r ) (53)
where gk ∈ SO(3) is an arbitrary rotation and ul ∈ C 2 (R1 ) are some real twice differentiable
functions. Neglecting questions of convergence for the moment, a specific solution is
constructed analogously to d’Alembert’s method of characteristics by choosing distinguished
functions pl for the functions ul and fitting them to the initially given pole density functions
by
X
P (hi , r ) = pl ((hi · g r )) (54)
l

where the pl belong to some dense specific subset of C 2 (R1 ), e.g. the set of polynomials.
Proposition. Let u : S 2 × S 2 7→ R be a twice continuously differentiable function,
u ∈ C 2 (S 2 × S 2 ), satisfying (1h − 1r )u(h, r ) = 0. If u reduces to a function of h · g0 r
for some arbitrary fixed g0 ∈ SO(3), then u is uniquely determined provided for some given
h0 ∈ S 2 it is known for all g0 r on a great circle containing h0 .

Proof. Any solution of the differential equation must be of the form (52) which according to
the assumptions reduces to u(h, r ) = u(h · g0 r ). For a given h0 and g0 r covering a great
circle containing h0 , t = h0 · g0 r covers the interval [−1, 1], and u is assumed known, i.e.
u(h0 · g0 r ) = v(t) for all r on that great circle. Eventually, u(h, r ) = u(h · g0 r ) = v(t). 
Considering the same assumptions with respect to spherical harmonics, the general
solution (43) simplifies to
X

u(h, r ) = λl Pl (h · g0 r ). (55)
l=0
With respect to practical texture analysis these assumptions appear rather artificial;
nevertheless, the approach by characteristics can be generalized.
Below it will be generally shown that a decomposition with an appropriate choice of
functions pl is equivalent to the series expansion into spherical harmonics. Moreover, if
the initially given pole density functions are expanded into such a series the solution of the
ultrahyperbolic differential equation specified by them is known. The crucial problems of
existence and uniqueness of a special solution will be discussed in a forthcoming paper; here
we provide some preliminary formulae.

4.4. Equivalent expansions of functions on the sphere


Any continuous (at least once) differentiable function on the sphere S 2 can be expanded into an
absolutely convergent series of spherical harmonics. It turns out that there exists an equivalent
decomposition, which is just derived by combination of specified spherical harmonics, but the
argument of these functions is the scalar product of h and r . This equivalent decomposition
provides the means to represent any pole density function in the form (54), cf [27, 28].
A function P ∈ C 2 (S 2 ) can be expanded into a series of spherical harmonics
X
∞ X
l
P (r ) = Flm Ylm (r ), (56)
l=0 m=−l
with harmonic coefficients
Z
Flm = P (r 0 )Ylm (r 0 ) dr 0 (57)
S2
1614 D I Nikolayev and H Schaeben

where r and r 0 are unit vectors of the sphere S 2 with spherical coordinates (ϕ, ϑ) and (ϕ 0 , ϑ 0 ),
respectively.
Alternatively, it is possible to represent a function P ∈ C 2 (S 2 ) as follows:
X∞
P (r ) = Xl (r ),
l=0
with functions Xl (r ) defined as
Z
2l + 1
Xl (r ) = P (r 0 )Pl (cos ν) dr 0 (58)
4 S2
where Pl denotes the Legendre polynomial of order l and ν denotes the angle enclosed by r
and r 0 , i.e.
cos ν = (r · r 0 ) = cos ϑ cos ϑ 0 + sin ϑ sin ϑ 0 cos(ϕ − ϕ 0 ). (59)
One can easily verify that
X
l
Xl (r ) = Flm Ylm (r ). (60)
m=−l

Next, we consider the series


X
∞ X l
P (r ) = 8m
l Pl (cos νm ), (61)
l=0 m=−l

where cos νm = (r · rm ) = cos ϑ cos ϑm + sin ϑ sin ϑm cos(ϕ − ϕm ). Then the problem is to
find 8m
l and rm such that (56) and (61) become identical.
According to the addition theorem for Legendre polynomials
X
l

Pl (cos νm ) = Ylk (r )(Ylk ) (rm ). (62)
k=−l

It is sufficient to consider representations of the decomposing functions Xl (r ) in terms of sums


of Legendre polynomials (62). Substituting (62) into (60) and rearranging the sums we obtain
X
l
8kl Ylm (rk ) = Flm . (63)
k=−l

Thus, if the coefficients 8kl and directions rk are known, equation (63) allows one to find the
coefficients Flm of the series expansion (56). In turn, suppose the directions rk are given;
then (63) is a system of linear algebraic equations for 8kl , with matrix {Ylm (rk )}. Multiplying
both sides of (63) by the complex conjugate matrix {(Ylm )∗ (rk )}, we obtain
X l X l Z
m ∗ m ∗
fl =
k
Fl (Yl ) (rk ) =
m
(Yl ) (rk ) P (r 0 )Ylm (r )0 dr 0
m=−l m=−l S2
Z
2k + 1
= P (r 0 )Pl (cos νk ) dr 0 (64)
4 S2
with
cos νk = (r 0 · rk ) = cos ϑ 0 cos ϑk + sin ϑ 0 sin ϑk cos(ϕ 0 − ϕk ). (65)
Applying equation (62) to the product on the left-hand side of equation (6) we finally get
X
l
8kl Pl ((rk · rm )) = flm . (66)
k=−l
Characteristics of the spherical ultrahyperbolic differential equation 1615

Another way to obtain equation (66) is to multiply both sides of equation (60) by
Pl ((rk · rm )) and integrate over the unit sphere.
The crucial problem now concerns the existence of a set of directions rk such that the
matrix Pl ((rk · rm )) of equation (66) is not degenerate.
Theorem 4.4. For each l ∈ N there exist 2l + 1 directions {rk } such that matrix Pl ((rk · rm ))
of equation (66) is not degenerate.

Proof. Suppose the opposite, or that the matrix of equation (66) is degenerate for any set
{rk } of 2l + 1 directions. Then it has linearly dependent rows, implying that some row can be
expressed as a linear combination of others for any direction rk on the sphere. But that means
linear dependence of spherical harmonics. This contradiction states the proof; see also [12],
pp 49–50. 
However, the choice of the set {rk } is not unique. Freedom in that choice allows us to
exploit additional knowledge of the function to be expanded. A preferential choice could be one
that turns the matrix of equation (66) into a special form. For example, choose rk = (ϑ0 , ϕk )
with ϕk = 2l+1
2πk
, k = 0, . . . , 2l, and an arbitrary angle ϑ0 . Then the matrix of equation (66)
becomes of Toeplitz form. The same result will be accomplished if we choose rk = (ϑk , ϕ0 )
with ϑk = 2l+1
πk
, k = 0, . . . , 2l, and an arbitrary angle ϕ0 .

4.5. Symmetrization
Crystal symmetry is the single predominant issue of texture analysis. Therefore, this section
is devoted to describing how to take crystal and sample symmetries into account. The
symmetrization is based on the linearity of the ultrahyperbolic differential equation.
Suppose, GB = {gBj , j = 1, . . . , NB } is the point symmetry group of a crystal,
and GA = {gAk , k = 1, . . . , NA } is the point symmetry group of the sample. Since
the ultrahyperbolic differential equation is satisfied identically for arbitrary functions of an
argument (h · g r ) for any g ∈ SO(3), it is also satisfied for arbitrary functions of argument
(gBj hi · ggAk r ) for all gBj and gAk in the case of crystal/sample symmetries, and moreover
u(xg ) = u(xgBj ggAk ).
This can be achieved if symmetrized polynomials Pl are used for the series expansion of a pole
density function according to
1 X NB XNA
P̃l (xg ) = Pl (xgBj ggAk ).
NB NA j =1 k=1

4.6. Algorithm to solve the ultrahyperbolic differential equation


According to (ht · g r ) = ((g t h)t · r ), we may interpret (r · rk ) as ((gkt h)t · r ). Then, it is
possible to summarize any method of solution as follows.
• Choice of the set of orientations {gk }.
That step seems hardest to formalize. The set {gk } could be chosen to diminish the required
computational efforts. From this point of view, a choice to obtain a Toeplitz matrix looks
promising. Alternatively, if it is possible to see peak or axial texture components in the
pole density functions, a more natural choice of the set {gk } could be one exploiting this
additional information. Such a choice could be done interactively with a computer.
• Determination of the set {rk } as induced by the set {gk }.
1616 D I Nikolayev and H Schaeben

• Series expansion of the pole density functions into symmetrized polynomials.


• The resulting coefficients are the coefficients of the required solution.
A specific and successful choice of the set {gk } is expected to depend on the specific
problem, particularly on assumptions which are reasonable with respect to this problem; at
this time it does not seem possible to give general rules.

4.7. Applications
Before discussing the existence and uniqueness of the solution of the ultrahyperbolic
differential equation it seems worthwhile to give an outlook of potential applications. The
most interesting for the practice of texture analysis include:
• defining a minimum set of experimentally accessible pole density functions required;
• defining an appropriate set of experimentally accessible pole density functions with respect
to
(1) numerical stability of the system
(2) experimental errors which depend on the crystal form
• completing incomplete pole density functions to normalize them;
• checking the compatibility of experimental pole density functions corresponding to
different crystal forms;
• calculating additional pole density functions of special crystal forms which cannot be
measured in diffraction experiments directly from the experimentally given ones;
• calculating the even portion of an orientation density function explaining the given pole
density functions.
The solutions of these particular problems will become straightforward once the existence
and uniqueness of a solution for given experimental intensity data P̃ (hq , rp ), rp ∈ D ⊂ S 2 ,
representing initial (or boundary) conditions is guaranteed, that is, once it has been clarified
that the given data can be uniquely represented in the form (54).

5. Conclusions

In this paper we have shown that the central mathematical relationship of texture goniometry
can be equivalently formulated in terms of the conventional integral equation representing a
hyperspherical analogue of the x-ray transform, and in terms of an ultrahyperbolic differential
equation which characterizes the range of the transform. Its characteristics have been
determined and geometrically interpreted as the familiar fibres of texture analysis. Thus, the
mathematical grounds for methods to solve the differential equation numerically are prepared
and new approaches to solve actual problems of quantitative texture analysis are opened.

Acknowledgments

The authors would like to thank Dr E T Quinto, Tufts University, Medford, MA, USA, for
substantial reviews and suggestions pointing to references including more general results than
ours and different proofs, and essentially contributing an outline of the proof of our main
result alternative to our initial one. HS would like to thank Alexander G Ramm, Kansas
State University, Manhattan, KS, USA, for a sympathetic discussion while on leave at the
University of Giessen, Germany, that helped to clarify the essential statements, and Gerald van
den Boogaart, Freiberg University of Mining and Technology, Germany, for the alternative
proof of the invertibility of the spherical x-ray transform using a result of [25].
Characteristics of the spherical ultrahyperbolic differential equation 1617

Appendix. Geometric representation of the path of integration and of characteristics

The set {g ∈ SO(3)|g r = h} may be represented as a great circle on the unit sphere S+3 ⊂ R4
by virtue of the Rodrigues representation q ∈ S+3 (‘quaternion’) of a proper rotation g ∈ SO(3);
for details of the Rodrigues representation the reader is referred to [1, 3, 23, 30].
For r 6= ±h let r · h = cos η and
h×r 1
n1 = = (h × r ) ∈ S 2 (67)
kh × r k sin η
then g1 = g1 (n1 , η) is the coordinate transformation according to the rotation g1 : KA 7→ KB
with the smallest angle of rotation such that r is mapped onto h, i.e. g1 r = h. If r = h, then
η = 0 and n1 = r .
Next let
h+r 1
n2 = = (h + r ) ∈ S 2 (68)
kh + r k 2 cos(η/2)
then g2 = g2 (n2 , π) is the coordinate transformation corresponding to a rotation with the
largest angle of rotation such that r is mapped onto h. If q1 , q2 ∈ S+3 are the unit vectors of
Rodrigues parameters representing the transformations (passive rotations) g1 , g2 , respectively,
   
n11 sin(η/2) n21
 n sin(η/2)  n 
q1 =  12 , q2 =  22  (69)
n13 sin(η/2) n23
cos(η/2) 0
then each rotation gt represented by
q (t) = q1 cos(t/2) + q2 sin(t/2), t ∈ [−π, π] (70)
maps r onto h.
The set
 
t t
Q(r , h) = q (t)|q (t) = q1 cos + q2 sin , t ∈ [−π, π] (71)
2 2
is easily identified as a great circle restricted to S+3 .
If r = −h, then the set {g ∈ SO(3)|g r = h} may be represented by the great circle
spanned by
   
q11 q21
q  q 
q1 =  12  , q2 =  22  (72)
q13 q23
0 0
with two linearly independent (q11 , q12 , q13 )t , (q21 , q22 , q23 )t ∈ S 3 orthogonal to r .
As q (t) varies within Q(r , h) according to equation (71), the corresponding angle of
rotation ω(t) varies according to
η t ω(t)
q4 (t) = cos cos = cos , t ∈ [−π, π]. (73)
2 2 2
Analogously, for r 6= h the set {g ∈ SO(3)|g r = −h} may be represented by the great
circle Q(r , −h) spanned by
   
−n11 sin π−η
2 n41
 −n12 sin π−η  n 
q3 =  2 
 −n13 sin π−η  , q4 =  42  (74)
2
n43
cos π−η2
0
1618 D I Nikolayev and H Schaeben

with
r−h 1
n4 = = (r − h) ∈ S 2 . (75)
kr − hk 2 sin(η/2)
If r = h, then the set {g ∈ SO(3)|g r = −h} may be represented by the great circle
spanned by
   
q31 q41
q  q 
q3 =  32  q4 =  42  (76)
q33 q43
0 0
with two linearly independent (q31 , q32 , q33 )t , (q41 , q42 , q43 )t ∈ S 2 orthogonal to r .
Due to the construction with ni · nj = δij , i, j = 1, 2, 4, and n1 × n2 = n4 ,
n2 × n4 = n1 , n4 × n1 = n2 , and further n1 · h = n1 · r = 0, n2 · h = n2 · r = cos(η/2),
n4 · h = −n4 · r = − sin(η/2), it should eventually be noted that the vectors qi , i = 1, . . . , 4,
are mutually orthonormal, qi · qj = δij , i, j = 1, . . . , 4.
The angle of any two points q ∈ Q(r , h) and q 0 ∈ Q(r , −h) is arccos q · q 0 = π/2, they
are referred to as being orthogonal; the orientation distance of the corresponding orientations
is 2 arccos q · q 0 = π. The orientation distance of an arbitrary g0 ∈ SO(3) represented by
q0 ∈ S+3 from the set Q(r , h) is inf t∈[−π,π ] 2 arccos q0 · q (t) = arccos h · g0 r , its orientation
distance from the set Q(r , −h) is arccos(−h · g0 r ) = π − arccos h · g0 r and
h · g0 r = (2ψ 2 − 1) cos η + 2ψ sin η(Ψ · n0 ) + 2(Ψ · r )(Ψ · h) (77)
with Ψ = and ψ = cos
sin( ω20 )n0 ω0
2
, where n0
and ω0 denote axis and angle of the orientation
g0 (cf [19]).
The set of all rotations g 0 with constant orientation distance h · g 0 r = h · g0 r from
the set {g ∈ SO(3)|g r = h} is the set of all rotations mapping r on the small circle
C(h, ) = {h0 ∈ S 2 |h · h0 = h · g0 r } with centre h and angle  = arccos h · g0 r .
Moreover, the set of all rotations q with constant orientation distance  from the set Q(r , h)
is given by
  

3 s s 
Q(r , h; ) = q (s, t) ∈ S+ q (s, t) = q1 cos + q2 sin cos
2 2 2
  
t t 
+ q3 cos + q4 sin sin , s, t ∈ [−π, π] . (78)
2 2 2
For any pair of parameters s 0 , t 0 ∈ [−π, π] with constant difference s 0 − t 0 = δ ∈
[−π, π ] q (s 0 , t 0 ) maps a given r onto the same h0 . If δ = 0, then h0 is the unit vector
0
with arccos h · h0 =  in the plane spanned by r and h, let it be denoted h0 ; if δ > 0
(δ < 0), then h0 is the vector of the small circle C(h, ) which results from a clockwise
0
(counterclockwise) rotation of h0 about h by an angle δ.

References

[1] Altmann S L 1986 Rotations, Quaternions, and Double Groups (Oxford: Oxford University Press)
[2] Ásgeirsson L 1936 Über eine Mittelwerteigenschaft von Lösungen homogener linearer partieller
Differentialgleichungen zweiter Ordnung mit konstanten Koeffizienten Ann. Math. 113 321–46
[3] Becker R and Panchanadeeswaran S 1989 Crystal rotations represented as Rodrigues vectors Textures
Microstruct. 10 167–94
[4] Bukhgeim A L 1988 Introduction into the Theory of Inverse Problems (Moscow: Nauka) in Russian
[5] Bunge H J 1965 Zur Darstellung allgemeiner Texturen Z. Metallkunde 56 872–4
[6] Bunge H J 1969 Mathematische Methoden der Texturanalyse (Berlin: Akademie)
Characteristics of the spherical ultrahyperbolic differential equation 1619

[7] Bunge H J 1982 Texture Analysis in Materials Science (London: Butterworths)


[8] Buerger M J 1971 Introduction to Crystal Geometry (New York: McGraw-Hill)
[9] Butzer P L and Nessel R J 1971 Fourier Analysis and Approximation (Boston, MA: Birkhauser)
[10] Courant R and Hilbert D 1953 Methods of Mathematical Physics vol 1 and 2 (London: Interscience)
[11] Erdélyi A (ed) 1953 Higher Transcendental Functions vol 2 (New York: McGraw-Hill)
[12] Freeden W, Gervens T and Schreiner M 1998 Constructive Approximation on the Sphere (With Applications to
Geomathematics) (Oxford: Clarendon)
[13] Funk P 1916 Über eine geometrische Anwendung der Abelschen Integralgleichung Math. Ann. 77 129–35
[14] Gel’fand I M, Minlos R A and Shapiro Z Ya 1963 Representations of the Rotation and Lorentz Groups and Their
Applications (Oxford: Pergamon)
[15] Harris G B 1952 Quantitative measurements of preferred orientation in rolled uranium bars Phil. Mag. 43 113–23
[16] Helgason S 1984 Groups and Geometric Analysis (New York: Academic)
[17] Ivanova T M and Savyolova T I 1994 Calculation of domains of dependence of pole figures for quartz Phys.
Solid Earth 29 505–9 (Engl. transl.) (Russian edn June 1993)
[18] John F 1938 The ultrahayperbolic differential equation with four independent variables Duke Math. J. 4 300–22
[19] Kunze K 1991 Zur quantitativen Texturanalyse von Gesteinen: Bestimmung, Interpretation und Simulation von
Quarzteilgefügen Diss. RWTH Aachen
[20] Matthies S 1979 On the reproducibility of the orientation distribution function of texture samples from pole
figures (ghost phenomena) Phys. Stat. Solidi b 92 K135–8
[21] Matthies S, Vinel G W and Helming K 1987 Standard Distributions in Texture Analysis vol 1 (Berlin: Akademie)
[22] Matthies S, Vinel G W and Helming K 1990 Standard Distributions in Texture Analysis vol 3 (Berlin: Akademie)
[23] Morawiec A and Pospiech J 1989 Some information on quaternions useful in texture calculations Textures
Microstruct. 10 211–6
[24] Müller C 1966 Spherical Harmonics (New York: Springer)
[25] Müller C 1998 Analysis of Spherical Symmetries in Euclidean Spaces (New York: Springer)
[26] Muller J, Esling C and Bunge H J 1981 An inversion formula expressing the texture function in terms of angular
distribution functions J. Physique 42 161–5
[27] Nikolayev D I 1994 Optimization algorithm for the Bunge/Roe method Phys. Solid Earth 29 518–22 (Engl.
transl.)
[28] Nikolayev D I 1994 Numerical economization of the series method Proc. 10th Int. Conf. on Textures of Materials
(Materials Science Forum vol 157–162) ed H J Bunge, pp 393–400
[29] Patch S K 1998 CT programs memo #98-05 Internal Report GE Corporate Research and Development
[30] Prentice M J 1986 Orientation statistics without parametric assumptions J. R. Statist. Soc. B 48 214–22
[31] Roe R J 1965 Description of crystallite orientation in polycrystalline materials: III. General solution to pole
figure inversion J. Appl. Phys. 36 2024–31
[32] Sansone G 1959 Orthogonal Functions (London: Interscience)
[33] Savelova T I 1982 Solution of one inverse diffraction problem Dokl. Akad. SSSR 266 590–3
[34] Savelova T I 1990 A numerical algorithm for solving the Cauchy characteristic problem for ultrahyperbolic
equations USSR Comput. Math. Math. Phys. 30 233–7
[35] Savyolova T I 1994 Inverse formulae for orientation distribution function Proc. 10th Int. Conf. on Textures of
Materials (Materials Science Forum vol 157–62) ed H J Bunge, pp 419–21
[36] Savyolova T I and Nikolayev D I 1993 Private communication
[37] Schneider R 1969 Functions on a sphere with vanishing integrals over certain subspheres J. Math. Anal. Appl.
26 381–4
[38] Seeley R T 1996 Spherical harmonics Am. Math. Mon. 78 115–21
[39] Szegö G 1939 Orthogonal Polynomials (4th edn 1975) (New York: American Mathematical Society)
[40] Vilenkin N J 1968 Special functions and the theory of group representations Am. Math. Soc. Transl. 22
[41] Wahba G 1981 Spline interpolation and smoothing on the sphere SIAM J. Sci. Statist. Comput. 2 5–16
Wahba G 1981 SIAM J. Sci. Statist. Comput. 3 385–6 (erratum)

You might also like