A Compact Symmetric Space: The Sphere
A Compact Symmetric Space: The Sphere
A Compact Symmetric Space: The Sphere
It is important that you establish a steady, rhythmic tempo while doing the exercises and
that you rest as little as possible between them. Also remember not to strain yourself.
From Jane Fondas Year of Fitness and Health/1984, Desk Diary, Simon and Schuster,
NY, 1983. Reprinted by permission.
107
108
group, SO(n), is the subgroup of matrices U in O(n) such that the determinant of
U is one. You can regard U in SO(n 1) as an element of SO(n) by forming
U0
0 1
in SO(n).
The group O(n) is a compact group and the sphere is a compact symmetric space,
making some of the analysis on it somewhat easier.
Exercise 2.1.1 (The Sphere as a Quotient or Homogeneous Space). Consider
the sphere Sn1 = {x Rn | x = 1}. Show that Sn1 can be identified with the
quotient space SO(n)/SO(n 1), using the preceding identification of SO(n 1) as
a subgroup of SO(n).
Hint. Map the coset gSO(n 1) to the vector gen for g in SO(n 1) and en =
t (0, . . . , 0, 1).
You may also want to read the discussion of the topology of spheres and
orthogonal groups in Chevalley [85, pp. 5267]. In particular, the fundamental (or
Poincare) group of SO(2) is isomorphic to Z, while that of SO(n), n 3, has order 2.
From now on we shall consider only the sphere S2 . See the references for the
general case. Now the sphere S2 is a differentiable manifold. This means that
locally it looks like two-dimensional Euclidean space. To make this precise, we use
the usual angular coordinates ( , ), 0 < < 2 , 0 < < , pictured in Fig. 2.1,
to parameterize S2 except for the semi-circle C through the poles and (1, 0, 0).
A similar coordinate patch can be constructed to cover the rest of the sphere. The
equations for the rectangular coordinates (x, y, z) of a point on S2 in terms of the
angular coordinates are
x = sin cos
y = sin sin
z = cos
(2.1)
109
More information on differentiable manif olds can be found in Yvonne ChoquetBruhat et al. [87], Helgason [276278], Loomis and Sternberg [428], Singer and
Thorpe [604], or Spivak [614], for example.
Before proceeding further with our discussion of spherical geometry, let us
review how to change variables in the Laplacian, using the method of Courant
and Hilbert [111, pp. 224225]. Assume that the substitution mapping (x, y, z) to
(u1 , u2 , u3 ) is differentiable with differentiable inverse. Let the Jacobian matrix of
the change of variables be
(x, y, z)
.
(u1 , u2 , u3 )
A=
(2.2)
dx
du1
ds2 = (dx dy dz) dy = (du1 du2 du3 ) t AA du2 .
du3
dz
Thus we obtain
ds2 =
gi j dui du j
where
G = t AA = (gi j )1i, j3 .
(2.4)
i, j=1
f f f
,
,
x y z
=
f f f
,
,
u1 u2 u3
A1 ,
to see that
f
x
2
+
f
y
2
+
f
z
2
=
i, j=1
gi j
f f
where G1 = (gi j )1i, j3 . (2.5)
ui u j
To see what happens to the Laplacian in the new coordinate system, one can use
the calculus of variations (see Courant and Hilbert [111, Chap. 4]). Suppose that f
minimizes the integral
J( f ) =
subject to the constraint
110
K( f ) =
f 2 dx dy dz = constant.
gi, j fui u j
|G| du1 du2 du3 ,
with
|G| = det G,
i, j=1
and
K( f ) =
f2
|G| du1 du2 du3 .
ui
i
ui |G| 2 gi,k uk .
i=1
(2.6)
k=1
There are other ways to obtain these formulas. See, for example, the following
references: Arfken [13, Chap. 2], Yvonne Choquet-Bruhat et al. [87, p. 307],
Churchill and Brown [89, pp. 1619], or Helgason [277, pp. 386387]).
Now we can show that the sphere is a Riemannian manifold, meaning that there
is a notion of arc length defined from an inner product on the tangent space to the
surface at each point. In order to obtain the arc length element on S2 , we shall use
the preceding discussion to write the Euclidean arc length element in spherical polar
coordinates. This is done in the next exercise.
Exercise 2.1.2 (Spherical Polar Coordinates). Spherical polar coordinates for a
point (x, y, z) in R3 are defined by
x = r sin cos , y = r sin sin , z = r cos ,
111
It follows from Exercise 2.1.2 that the element of arc length on the unit sphere
S2 is
ds2 = d 2 + sin2 d 2 .
(2.7)
This element of arc length is invariant under O(3) as is the corresponding volume
or area element:
d = sin d d .
(2.8)
1
sin
sin
+
1 2
sin 2
.
(2.9)
Exercise 2.1.3. Show that the Euclidean Laplacian f = fxx + fyy + fzz is invariant
under O(3) (i.e., that commutes with rotation). Deduce that the spherical
Laplacian (2.9) is also invariant under O(3).
Hint. Dym and McKean [147, p. 243], have a proof using the Fourier transform.
Because the sphere is a Riemannian manifold, we can consider geodesics on the
sphere. The geodesic through two points on the sphere is the curve through these
two points that minimizes distance. Airplane pilots know that geodesics in S2 are
great circles; i.e., the intersection of S2 with a plane through the origin and the two
given points on S2 . To see this, suppose that you are given points p and q on S2 . You
can rotate S2 so that both points p and q have -coordinate in (2.1) equal to zero.
See Exercise 2.1.11 if you do not believe this. Then the distance between p and q
on S2 is
q
q
d 2 + sin2 d 2
d = (q) (p).
p
Thus the great circle route minimizes distance on the sphere (assuming that you go
in the proper direction around the circle).
112
113
dn
2
(x 1)n ,
n
dx
for n = 0, 1, 2, . . . ,
(2.11)
d m Pn (x)
.
dxm
(2.12)
(n m)! m
P (x).
(n + m)! n
+1
1
f (x)g(x) 1 x2 dx.
This means that you must show that fkm is orthogonal to xr when r = 0, 1, . . . , k 1.
Integration by parts, using the case m = 0, will do the trick.
114
(n + m + 1)
(n + 1)
0
x+
x2 1cos n cos(m ) d
for Re x > 0, m = 0, 1, 2, . . ..
Hint. See Courant and Hilbert [111, p. 505].
Exercise 2.1.7 has been generalized to symmetric spaces by Harish-Chandra.
This will be discussed in later chapters.
Theorem 2.1.1 (Spherical Harmonics).
(a) The only eigenvalues of the spherical Laplacian defined in (2.10)
are = n(n + 1), n = 0, 1, 2, . . .. The vector space of eigenfunctions of
corresponding to the eigenvalue = n(n + 1) has dimension 2n + 1.
A complete orthogonal set of eigenfunctions of is
Y ( , ) = exp(im )Pnm (cos )
for n = 0, 1, 2, . . . , |m| n.
(2.13)
Here Pnm is the associated Legendre function (2.12) and the coordinates ( , )
are defined by (2.1). We call cY , with Y given by formula (2.13), a (surface)
spherical harmonic of degree n and order m, where c is a normalizing
constant.
(b) Let f (x, y, z) = rnY ( , ), using spherical polar coordinates (r, , ) corresponding to the rectangular coordinates (x, y, z) in Exercise 2.1.2. Then Y is
a surface spherical harmonic satisfying
Y = n(n + 1)Y
if and only f (x, y, z) is a homogeneous harmonic polynomial of degree n. When
we say that f is harmonic, we mean that f satisfies Laplaces equation
f = fxx + fyy + fzz = 0.
Exercise 2.1.8. Prove part (a) of Theorem 2.1.1, using separation of variables on
Y = Y and Exercise 2.1.6.
Note. There are other proofs that { = n(n + 1), n = 0, 1, 2, . . .} are the only
possible eigenvalues of on S2 . For example, see Dym and McKean [147, pp.
252253], Kirillov [353, pp. 271274], or Van der Waerden [691, p. 21].
Exercise 2.1.9. Use separation of variables on f = 0 to prove part (b) of
Theorem 2.1.1. Note that Exercise 2.1.2 says that if f (x, y, z) = R(r)Y ( , ), then
115
f =
1 2
R
(r R (r)) Y + 2 Y
r2
r
where
d
.
dr
Kg K,
with
cos
g = 0
sin
0 sin
1
0 ,
0 cos
0 < .
116
Y00 = 14
Y 1 = 3 sin exp (i )
1
8
3
Y0 =
cos
1
4
15
2
5 3
Y0 =
cos2 12
2
4
2
105
Y3 = 32
sin2 cos exp (2i )
1
21
2
7 5
Y0 =
cos3 3 cos
3
When do two of these cosets coincide? The corresponding points to gK on the sphere
are ge3 , where e3 = t (0, 0, 1). So if
cos sin 0
k = sin cos 0 ,
0
0 1
we find that k g e3 = t (cos sin , sin sin , cos ). Thus a function on K\G/K
is a function only of the angle .
Let A = {g , |0 }, with g from Exercise 2.1.11. Then the decomposition
SO(3) = KAK, which follows from the preceding exercise, is the Euler angle
decomposition of SO(3). See Volume II [667] for a generalization of this decomposition and its application to harmonic analysis on general symmetric spaces.
Exercise 2.1.12. Check Table 2.1 which lists the first few surface spherical harmonics. The table uses the CondonShortley [101] convention.
We can use Mathematica to visualize spherical harmonics using density plots on
7
the sphere. See Fig. 2.3 for two such pictures. On the left is that of ReY14
( , )
obtained using the Mathematica command:
ParametricPlot3D[ {Cos[ ] Sin[ ], Sin[ ] Sin[ ], Cos[ ]}, { , 0, 2 }, { , 0, },
PlotPoints ->300,Mesh->None,ColorFunction->Function[ {x, y, z, , },
Hue[Re[SphericalHarmonicY[14,7, , ]]]],ColorFunctionScaling->
False].
7 ( , )+2ReY 3 ( , ) obtained using a similar
On the right is that of ReY14
7
command.
117
These should be compared with analogous plots for the analogous functions on
a unit square in the plane in Fig. 1.4, which we considered in Sect. 1.3.
Exercise 2.1.13. (a) Make some similar plots to those in Fig. 2.3 using the Mathematica command
Re[SphericalHarmonicY [m, n, , ]].
(b) Solve the wave equation on the sphere with initial condition modelling what
would happen if the sphere were struck with a hammer at a point. Then make
a Mathematica movie of your solution.
(c) Consider the effect of forcing a sphere to vibrate at some frequency near an
eigenfrequency of .
Corollary 2.1.1 (Harmonic Analysis on the Sphere). Every function f : S2 C
with continuous second-order derivatives can be expanded in an absolutely and
uniformly convergent series of spherical harmonics. If we take Ynm , |m| n, to be
an orthonormal basis for L2 (S2 , d ), with d as in Exercise 2.1.2, then
f ( , ) =
n0 |m|n
f(n, m)Ynm ( , ),
where
f(n, m) =
S2
f Ynm d .
118
2
e2
+
2m
r
= E,
r = (x2 + y2 + z2 )1/2 .
(2.14)
119
of the system if me is the mass of the electron and m p the mass of the proton. The
eigenvalue E is the energy level of the system. Some references for these matters
are Biedenharn and Louck [39], Castellan [80], Condon and Odabasi [100], Courant
and Hilbert [111, pp. 341343], Eyring, Walter, and Kimball [170], Mackey [442,
pp. 159271], Messiah [465, especially pp. 362 and 412], Sommerfeld [610, pp.
200206], B. Thaller [676], Van der Waerden [691, Chap. 1], Dorothy Wallace and
J. BelBruno [713], Weyl [731, pp. 4170], and Wigner [738].
The eigenvalues E that lie in the discrete spectrum of the Schrodinger operator in
equation (2.14) will be the only ones of interest for our discussion. Such eigenvalues
are negative (in fact, there is also a continuous spectrum of positive real numbers).
Physicists interpret the values of E that lie in the discrete spectrum as energy levels
of bound states of hydrogen. Spectroscopists going back to Balmer in the 1880s
have found various series of lines in the spectrum of hydrogen corresponding to
these energy levels. We can obtain some understanding of these spectral lines by
finding the discrete spectrum of the Schrodinger operator very explicitly.
We shall use spherical harmonics to separate variables via (x, y, z) =
w(r)Y ( , ) in equation (2.14) for the purpose of obtaining the discrete spectrum
eigenvalues E. Then Exercise 2.1.2 shows that (2.14) becomes
Y
(r2 w ) + (2mr2 /2 )(E + e2/r)w
=
,
w
Y
where
d
.
dr
w(r) = rn exp(cr/(2k))Lk+n
(cr/k),
me4
.
22 k2
(2.15)
k1
(2n + 1)
n=0
(2.16)
120
= Rk2 ,
where R is the Rydberg constant, R = 2 2 me4 h3 . When an atom moves from initial
state 1 with principal quantum number k1 to final state 2 with principal quantum
number k2 , the frequency of the associated spectral line is
= R(k22 k12 ).
The series of spectral lines of hydrogen observed by Balmer in 1885 has k2 = 2.
Thus the Balmer series lines have frequencies
= R(22 k2 ),
k = 3, 4, 5, . . . .
These lines are in the visible range. Lyman found a series of spectral lines in the
ultraviolet range in 1909 with k2 = 1. The series with k2 = 3, 4, are in the infrared
range. The Balmer series is not obtained in the laboratory because it requires the
temperature of stellar atmospheres.
The theory just described does not account for the fine structure of the spectrum.
Relativistic effects and spin must be considered. Even then, agreement between
theory and experiment is not perfect. See Messiah [465, Vol. II, pp. 930933], for a
discussion of the Lamb shift.
There is an additional degeneracy of (n + 1)2 rather than 2n + 1, which was
explained by Fock, who showed that if one formulates the Schrodinger equation
for the hydrogen atom as an integral equation, then it is also invariant under SO(4).
See Louck and Metropolis [430] for a discussion of a related problem which was
motivated by Focks result.
Spectroscopists daily analyze all sorts of materialsatoms, molecules, crystals,
solids, gases. But, of course, the theory becomes much more complicated for the
many-body problem. Group theory helps, as many of the references mentioned at
the beginning of this discussion show.
121
d p z p
(e z ) = the pth Laguerre polynomial.
dz p
(2n+1)
(cr/k) satisfies
e2
E+
w = n(n + 1)w,
r
for E given by equation (2.15), k Z, k > n, c = 2me2 /(2 ). Then show that we
have found the only continuous bounded eigenfunctions w(r) for the differential
operator involved here.
Hint. See Arfken [13, Chap. 13], or Courant and Hilbert [111, p. 330].
We have found a wave function solving equation (2.14) of the form
(2n+1)
(cr/k),
(b)
where Ynm denotes a spherical harmonic as in Theorem 2.1.1 and La is the bth
derivative of the Laguerre polynomial La . Then | |2 can be interpreted as the
probability density of the electron for the various states of the hydrogen atom.
Fig. 2.4 from White [733, p. 71], illustrates the first few states.
122
Fig. 2.4 Photographs of the electron cloud for various states of the hydrogen atom as made from
a spinning mechanical model. The probability-density distribution is symmetrical about the
-axis, which is vertical and in the plane of paper. (From H.E. White, [733, p. 71]. Reprinted by
permission of McGraw-Hill)
123
(r, , ) = R
n0
|m|n
1
1 a2n+1
R n+1
2n+1 r n
a
f(n, m)Ynm ( , ),
r
R
S2
f Ynm d .
Fig. 2.5 Open magnetic structures on the sun. (From Altschulers article in Herman [290, p. 125]. Reprinted by permission of Springer-Verlag)
124
2 A Compact Symmetric Space: The Sphere
125
correspond to the fields. There are many books on the subject. We leave it to you
to Google them, or see Wikipedia. There is also the Grand Unified Theory, or GUT.
This involves SU(5), SO(10) and even the exceptional Lie group E8 . Again there is
much on the Internet including some beautiful movies related to E8 .
Hecke and others have made use of these higher-dimensional spherical harmonics in number theory and the kinetic theory of gases (see Hecke [258, pp. 361373,
849854]) and Ogg [505, Chap. VI, p. 6]. We shall see in the next section that
spherical harmonics also give a new description of harmonic analysis on Rn .
126
m = ea (x) = exp t ax , for x Rm | a Rm
R
(2.17)
127
(a) Show that two finite-dimensional representations (T1 , H) and (T2 , H) are equivalent if and only if the matrix of T1 is obtained from the matrix of T2 by changing
the basis of H that is used.
(b) Show that a finite-dimensional representation (T, H) of G is completely
reducible if and only if you can find a basis of H which puts the matrix of
T in block diagonal form:
T1
.. . .
. .
0
..
.
0 Tn
so that the representation determined by the matrix Ti is irreducible for each
i = 1, . . . , n.
We will often need to be able to integrate functions on topological groups. The
integral involved is an analogue of the Lebesgue integral on Rn and is supposed to
come from a countably additive positive measure on the Borel sets in the group.
Such an integral on a topological group G is right invariant if
G
f (x) dx =
G
f (xa) dx
for all a in G.
The invariant integral on a compact Lie group was used by Hurwitz, Schur, and
Weyl beginning in the 1890s. Haar proved the existence of a invariant integral on
any locally compact topological group in 1933 and thus the integral is called the
Haar integral. If G = Rm , under addition, then the Haar integral is the standard
Lebesgue integral. Proofs of the existence of the Haar integral may be found in
Helgason [277, p. 365], Lang [387], Pontryagin [518], or Weil [725], for example.
The invariant integral can be used to show that any irreducible unitary representation of a compact group must be finite-dimensional (see Kirillov [353, p. 135]),
for example. And you will need the invariant integral to do Exercise 2.1.20. In fact,
integration on topological groups will be a necessary tool for the rest of this book.
Of course, our groups will be matrix groups and thus the Haar integral for these
groups is not so mysterious.
The right-invariant Haar measure dx is unique up to a positive constant multiple.
This means that for each g in G, there is a positive constant (g) (not to be confused
with the like-named distribution) defined by
f (gx) dx = (g)
f (x) dx.
128
f (x) (x) dx =
f (x1 ) dx.
The group G is said to be unimodular when (g) = 1 for all g in G. This means
that right-invariant Haar integrals are also left invariant. Many groups that we shall
consider are unimodular. In particular, all compact or abelian groups are unimodular.
Most of the Lie groups that we shall consider in this book are what is called
semisimple; for example, the special linear group of all n n real matrices of
determinant one. It can be shown that all semisimple Lie groups are unimodular (see
Helgason [277, p. 366]). However, not all groups are unimodular. For example, the
group of upper triangular matrices with positive diagonal entries is not unimodular
if n 3 (Exercise).
Exercise 2.1.19. Show that any finite-dimensional unitary representation is completely reducible and that its decomposition into irreducible representations is
unique up to equivalence.
Exercise 2.1.20 (The Haar Integral on a Compact Group).
(a) Show that any compact group is unimodular; i.e., right Haar measure = left
Haar measure.
(b) Use Haar measure to show that any representation of a compact group is
equivalent to a unitary representation.
Suppose that (T1 , H1 ) and (T2 , H2 ) are representations of G. Define the tensor
product representation (T1 T2 , H1 H2) by
(T1 T2 )(g)(v1 v2 ) = (T1 (g)v1 ) (T2(g)v2 ) for g G, vi Hi , i = 1, 2.
See the references mentioned above for more information on tensor products. The
decomposition of tensor products of representations into direct sums of irreducible
representations has much importance for quantum mechanics. For example, if
one ignores the interaction between two electrons in the field of a positive nucleus,
the Schrodinger operator of the system has eigenfunctions which are products 1 2
of eigenfunctions i corresponding to representations Ti of O(3), i = 1, 2. And the
product 1 2 corresponds to the representation T1 T2 . The ClebschGordon series
breaks T1 T2 up into its irreducible components. Thus one can conclude what sort
of spectral lines should occur for such a situation. See the references mentioned in
Sect. 2.1.3 on quantum mechanics for more details.
There are many other useful topics in representation theory such as induced
representations, Frobenius reciprocity, Cartans theorem on the highest weight,
and Weyls character formula. See the references on group representations for a
discussion of these matters.
The systematic study of group representations of finite groups began in the
1890s with work of Frobenius, Schur, and others. In the 1920s, Weyl obtained the
irreducible representations of the compact simple Lie groups such as G = SO(3) (see
Weyl [730, Vol. II, pp. 543647]). To do this, Weyl used his formula for the character
129
G
G
When G is a tame unimodular Lie group the abstract Plancherel theorem (or
Fourier inversion theorem) says there is a measure d on G called the Plancherel
measure such that
f (e) =
Tr T ( f ) d (T ),
130
|k|n
(a) Justify this formula for An (g) and then show that An (g) defines a representation
of SO(3).
(b) Show that An (g) is a unitary representation using
S2
S2
f (g) =
n=0 (2n + 1) Trace [ f (n)An (g)], where
f(n) =
SO(3)
f (g)
tA
(2.18)
This is proved, for example, in Dym and McKean [147, pp. 256261]. Vilenkin
[704, pp. 440457], examines the representations of SO(n) defined in an analogous
way to that used for n = 3 in Exercise 2.1.21 and shows that if n > 3, then
these representations do not exhaust all of the irreducible unitary representations
of SO(n). Only the class-one representations are obtained in this way. A class-one
representation (A, H) of G has a vector v in H such that A(k)v = v for all k in K.
Here G = SO(n) and K = SO(n 1) (embedded in G as in Exercise 2.1.1).
Formula (2.18) giving Fourier analysis on the group SO(3) is a special case of the
PeterWeyl theorem for any compact group (see Weyl [730,
+ Vol. III, pp. 5875],
,
Pontryagin [518], or Weil [725]). The theorem says that if T = (mi j ), A
is a complete system
of irreducible unitary representations of a compact topological
131
adjoint operators (Theorem 1.3.7 of Sect. 1.3). The compact self-adjoint operator
used in the proof of the PeterWeyl theorem is the convolution operator T f = g f ,
where g is a fixed nonzero continuous function on G such that g(x) = g(x1 )
for all x in G. The convolution is taken with respect to Haar measure on G. The
eigenfunctions for a fixed eigenvalue of T form a finite-dimensional vector space on
which G operates by sending f (x) to f (xa) for a G. This gives a representation of
G. The eigenfunctions for T form a complete orthonormal set in L2 (G) and thus so
do the matrix entries of the unitary irreducible representations of G. It follows that
the analogue of formula (2.18) replaces An by the irreducible unitary representations
of G and (2n + 1) by the degree of the irreducible representation.
The fact that special functions come from representations often leads to an easier
understanding of the myriads of formulas listed in books such as Erdelyi et al. [164,
165].
Exercise 2.1.22 (Addition Formulas for Matrix Entries). Suppose that (T, H) is
a finite-dimensional representation of G and that e1 , . . . ,en is an orthonormal basis
of H. Let the i, j th matrix entry of T be mi j (g) = (T (g)ei , e j ) for g in G. Here (, )
denotes the inner product on H. Show that T (gh) = T (g)T (h) implies the following
addition formula for the matrix entries:
mi j (gh) =
k=1
|m|n
2n + 1 t
Pn ( xy).
4
Note that both sides are unchanged if you replace x by gx and y by gy for g in
SO(3). So the right and left sides of the equation can only differ by a constant. Find
the constant by setting x = y and integrating over S2 .
132
(see Gelbart [200], Godement [213], Macdonald [440], Tamagawa [656]). This
should not, however, cause us to forget the differential equations and the contact
with applied mathematics.
The following theorem was proved in 1916 by Funk [187] and in 1918 by Hecke
[258, pp. 208214].
Theorem 2.1.2 (The FunkHecke Theorem). Suppose that Yn is a spherical
harmonic of degree n. Let f : [1, +1] C be continuous. Then
S2
+1
1
Proof. Since every continuous function on [1, +1] can be uniformly approximated
by zonal spherical functions Pn (x), it suffices to do the following exercises.
Exercise 2.1.24 (Properties of Spherical Harmonics).
(a) Show that
S2
Ym (x)Pn (t xy) dx =
4
Yn (y)mn .
2n + 1
Here mn = 0 if m = n and mn = 1 if m = n.
Hint. Use the addition formula of Exercise 2.1.23.
(b) Prove that
+1
1
Pn (x)Pm (x) dx =
2
mn .
2n + 1
K
Y (xky) dk = Y (x)
K
Y (ky) dk
for all x, y in G,
(2.19)
P(I)
K
(2.20)
133
SO(3)
f (u)g(xu1 ) du
for x in SO(3).
Define the Fourier transform f(n) for n = 0, 1, . . . as in formula (2.18). Show that
( f g)(n) = f(n)g(n)
for all n = 0, 1, 2, . . . .
Hint. Use the fact that An is a representation.
Exercise 2.1.27 (The Fundamental Solution to the Heat Equation on the Sphere
S2 R3 ). Given the initial heat distribution f ( ), solve the following initial value
problem:
u( , ,t) = ut ,
u( , , 0) = f ( ),
t > 0.
n>0
134
reasoning in a certain sense, the reasoning is far from rigorous. Approximations are
introduced at many steps with only intuition as a guide to the error involved. We do not
know of a single instance in which a tomographic algorithm has been justified in a truly
rigorous sense. Thus, in contrast to some other workers in this field, we do not feel that one
derivation is more rigorous than another, whether it is based on Radons inversion formula,
the Fourier inversion formula or any other foundation.
From Shepp and Kruskal [588, p. 421].
(k + 1)(k + + 1) ,
k=0
y + (1/z)y + 1 ( /z)2 y = 0.
(b) Show that J (z) is represented by the integral formula:
J (z) =
(z/2)
(1/2)( + 1/2)
+1
1
135
2 n+ 1
z 2
1 d
z dz
n
sin z
, n = 0, 1, 2, . . . .
z
J (z) (1
+ )
(z/2) , as z 0;
J (z) 2z cos z
2 4 , as z .
(b) Show that Jn (z) = (1)n Jn (z), when n = 0, 1, 2, . . .. Then show that if is not
an integer, J and J are linearly independent.
(c) Functional Equation. Prove that J (z) = exp( i)J (z).
(d) Addition Formula. Set R = (r12 + r22 2r1 r2 cos )1/2 and prove that
J 1 (R) =
2
1
1
2R
|g(r)|r2 dr < .
Then the Euclidean Fourier transform as defined in Sect. 1.2 of f (assuming that f
is nice enough for the transform to exist), is given by
2
f(ru) = n Y (u)
i r
sR+
Proof. Let dv denote the element of surface area on S2 . Then, using Theorem 2.1.2
of Sect. 2.1, we have
f(ru) =
sR+ vS2
= 2 Y (u)
sR+
g(s)s2
+1
1
1
2n n!
dn
dxn
136
+1
1
(i)n
exp(2 irst)Pn(t) dt = Jn+ 1 (2 rs).
2
rs
Use integration by parts and Exercise 2.2.2. Then complete the proof of
Theorem 2.2.1.
How does the formula of Bochner and Hecke fit into the general scheme of
harmonic analysis on symmetric spaces? Why did the Bessel functions suddenly
appear out of the blue? To understand this, one must view R3 as a symmetric space
rather than just as an additive group. We explained this at the end of Chap. 1. It
leads one to think of the full group of isometries of R3 , namely, the Euclidean
group M(3, R) of rigid motions of R3 . If we consider only the orientation-preserving
motions G, then G is the semidirect product of the groups T and K, where T consists
of translations and is isomorphic to R3 ,
and K= SO(3). One can view M(3, R) as
Ab
a group of 4 4 matrices of the form
, A SO(3), b R3 . The matrix
01
Ab
sends x R3 to Ax + b. Clearly one can identify R3 with G/K by mapping
01
a coset gK, g G, to g0 = the result of applying the motion g to the origin 0 in R3 .
Exercise 2.2.5. Do an analogue of the BochnerHecke theorem for R2 for rotationinvariant functions f on R2 .
Answer. For nice rotation-invariant functions f on R2 , one has
f( u) = 2
f (r)J0 (2 r)rdr.
137
harmonics of degree n (as in Theorem 2.1.1 of Sect. 2.1). Then any f in L2 (R3 )
has a Fourier expansion (converging in the L2 -norm) of the form
f (x) =
n0 |m|n t>0
where
f(n, m,t) =
R3
Exercise 2.2.7. (a) Prove formula (2.21) by writing down the inversion formula
for the ordinary Fourier transform (from Theorem 1.2.1 of Sect. 1.2) in spherical
polar coordinates. Then use Theorem 2.2.1 to evaluate the inner Fourier
transform. Finally, Theorem 2.1.2 of Sect. 2.1 and Exercise 2.2.4 should be used
to complete the proof.
(b) Show that formula (2.21) is equivalent by change of variables to a special case
of Hankels inversion formula: (assuming > 12 and r > 0) :
f (r) =
0
yJ (yr)
0
(2.22)
Hankels inversion formula (2.22) gives the spectral resolution of the singular
SturmLiouville operator
(L f ) (x) =
2
1
f
(x f ) +
x
x
for x (0, )
(2.23)
This spectral resolution can be derived from a formula of Stieltjes, Stone, Kodaira,
and Titchmarsh, which is itself a corollary of the Von Neumann Spectral Theorem
for unbounded operators. Let us summarize the theory briefly, following Lang [387],
Reed and Simon [540, Chap. 7], Stakgold [618, 619], and some unpublished notes
of J. Korevaar. More details can be found in these references as well as Dunford and
Schwartz [145], Gelfand and Vilenkin [207], Levitan and Sargsjan [415], Maurin
[459], Naimark [490], Titchmarsh [679], Weyl [730, Vol. I, pp. 195297], and
Yosida [748]. In particular, Dunford and Schwartz [145, Vol. II, pp. 13331392,
15321533], provides an extremely careful discussion of spectral theory for singular
differential operators, including the Bessel operator (2.23).
138
dE ,
T=
dE ,
(2.24)
for some family of projection-valued measures dE (cf. Reed and Simon [540,
Chap. 7]). The integrals are over the spectrum of T , which is real, because T
is
self-adjoint. The spectral theorem implies that for polynomials p( ), p(T ) =
p( )dE . It is possible to define f (T ) for continuous functions f on the spectrum
of T and then this same formula holds upon replacing p by f .
In order to obtain the formula of Stieltjes, Stone, Kodaira, and Titchmarsh one
must be able to compute the Greens function or resolvent kernel G( ; x, y)
defined by
(T I)1v(x) =
G( ; x, y)v(y) dy.
(2.25)
+
( )1g( )d ,
lim 0+ ( f +
(c + i ) f (c i ))
= lim 0+
1 +
( c)2 + 2
g( ) d
= g(c).
This is Poissons integral formula for a half-plane (see Ahlfors [3, p. 171]). It follows
that the spectral measure for the operator T is given by the formula of Stieltjes,
Stone, Kodaira, and Titchmarsh:
2 i
139
dE
= (T ( + i0)I)1 (T ( i0)I)1 ,
d
(2.26)
(p f ) + q f ,
w
a < x < b.
(2.27)
Here singular means that either the interval is infinite or the functions w(x) or
q(x) blow up, or p(x) vanishes at some point in [a, b]. The Hilbert space associated
to (2.27) is L2 ([a, b], w) consisting of Lebesgue measurable functions f on [a, b]
such that
b
a
b
a
f (x)g(x)w(x) dx.
If one makes the correct assumptions about p, q, w and if one imposes the correct sort
of boundary conditions at a and b, the operator L will be self-adjoint. In particular,
p, q, w should be real and w should be positive.
The Greens function G( ; x, y) of formula (2.24) must satisfy
p
G( ; x, y) + (q w)G( ; x, y) = (x y).
x
Pick c in (a, b). We want to choose two linearly independent solutions and
of Lu = u, as follows. Suppose that solves Lu = u and in addition lies in
L2 ([a, c], w). Let solve Lu = u and L2 ([c, b], w). Then
G( ; x, y) =
and
1
c
(2.28)
140
(x) (x)
.
c =
p (x) p (x)
Example 2.2.1 (Spectral Measure for the Hankel Transform Using the Kodaira
Titchmarsh Formula). Consider the SturmLiouville operator given by formula
(2.23). Then formula (2.28) becomes
i J ( 1/2 x)H(1) ( 1/2 y), 0 < x < y <
G( ; x, y) =
2 J ( 1/2 y)H(1) ( 1/2 x), 0 < y < x < .
(1)
Here J is the Bessel function of the first kind from Exercise 2.2.2 and H is the
Hankel function (or Bessel function of the third kind) defined by
H1 (z) =
J (x) 2x cos x
as
2 4 ,
(1)
H (x) i 2x ( ),
as
(1)
H (x) 2x exp i x
, as
2 4
x 0,
x ,
x 0,
(2.29)
x .
(1)
Note that if 0 < < 1, both J ( 1/2 x) and H ( 1/2 x) are in L2 ([0, 1], x). One says
that 0 is then in the limit-circle case in Weyls theory (cf. Stakgold [618,619]). If
(1)
1, then J ( 1/2 x) is in L2 ([0, 1], x) and H ( 1/2 x)
/ L2 ([0, 1], x). One says that 0
(1)
1/2
/ L2 ([1, ], x) and H ( 1/2 x)
is then in the limit-point case. Note that J ( x)
2
L ([1, ], x) so that infinity is always in the limit-point case.
To compute the jump of G( ; x, y) required by formula (2.26), assume that x < y.
Then for > 0,
G( + i0;
x, y) G( i0; x, y)
=
i
2
(1)
(1)
J ( 1/2 x)H ( 1/2 y) J ( 1/2 x)H ( 1/2 y) .
141
(1)
(1)
H (x) = exp( i) H (x) 2J (x) .
These formulas imply
G( + i0; x, y) G( i0; x, y) = i J ( 1/2 x)J ( 1/2 y).
Thus we obtain from (2.24) and (2.26),
(x y) 1
=
x
2
0
0
J ( x)J ( y) d .
cm J ( m x/a),
0 x a,
> 1,
(2.30)
m=1
a
0
0 < x < .
142
Here > 0 and is the eigenvalue. Show that the Greens function is
G( ; x, y) =
where I and K are the Bessel functions of imaginary argument. The definitions
are
I (z) = exp( i/2)J (z exp( i/2)),
K (z) =
i
(1)
exp( i/2)H (z exp( i/2)),
2
See Chap. 3 and Lebedev [401] for more information on these functions.
Show that K (x) = K (x) and I (x) I (x) = 2 sin( )K (x)/ . Then prove
the KontorovichLebedev inversion formula:
x (x y) =
2
2
Ki ( x)Ki ( y) sinh ( ) d .
t xu=k
xR2
f (x) ds,
(2.31)
where the integral is over a line in the plane and ds is the element of arc length. The
inversion formula for this transform goes back to Radon [528] in 1917. It says that
f (x) =
1
d
q>0 q
1
2
u S1
R f (q + t xu, u) du .
(2.32)
143
S1 one year earlier than Radon proved his result. Helgason [279] shows that a vast
generalization of this theory is possible, viewing the Radon inversion formula as
involving two dual integrationsone over the points in a hyperplane and the other
over the hyperplanes through a point. The Radon transform appears in so many
applications to signal processing that Matlabs Signal Processing Toolbox has a
Radon transform command. We look at finite analogues of the Radon transform
in [668, pp. 7172]
The next sequence of exercises presents a derivation of Radons inversion
formula.
Exercise 2.2.9. (a) Suppose that f : R2 C is a Schwartz function. Show that the
Fourier inversion formula can be written in the form
f (x) =
uS1
E f ( t ux, u) du,
where
E f (t, u) =
1
2
rR kR
In the first formula du denotes the angle measure on the unit circle S1 . The
Radon transform R f (k, u) is defined by (2.31).
(b) Let abs(r) = |r|, r R. Show that abs(r) is not the Fourier transform of a
Lebesgue integrable function.
were a function,
Hint. Recall the RiemannLebesgue lemma. Note that if abs(r)
then you could write:
E f (t, u) =
1
2
kR
abs(t
k)R f (k, u) dk.
f (x) dx = lim
|x|>
f (x) dx .
Define
(x1 , ) = PV
x1 (x) dx
and
2
(x , ) = PV
x2 ( (x) (0)) dx
144
for test functions as in Sect. 1.1. Prove that in the sense of distributions (see
Sect. 1.1) the following formulas are valid:
(a)
(b)
(c)
(d)
x1 = (log |x|) .
x2 = (x1 ) .
1 = i sgn(x), where sgn(x) = abs(x)/x.
x
= (2 2)1 x2 .
abs(x)
Hint. See Vladimirov [706, pp. 75, 86, 134], or Bracewell [61, p. 130].
Exercise 2.2.11. Derive Radons inversion formula from Exercises 2.2.9
and 2.2.10, using properties of Fourier transforms of distributions.
Note. The Hilbert transform of a function is
H f (x) =
+
f (t)
(x t)
dt =
1
f (x1 ).
Compare the decay at infinity with that of |x|2 from part (d) of Exercise 2.2.10.
Note. You can write the Fourier transform in the plane as a Bessel transform, much
as we did in Theorem 2.2.1 for the Fourier transform in 3-space. See Exercise 2.2.5.
Some History. In 1956, R. Bracewell used a method analogous to the Radon
transform to study solar radiation. In the 1960s and 1970s CT scanners for medicine
were developed independently by A. M. Cormack and G. N. Hounsfield. They were
not based on the Radon transform. The Hounsfield algorithm was used in the first
commercial CAT scanner made by EMI Central Research Labs. in the UK. The first
patient brain scan was done in 1971. A. M. Cormack and G. N. Hounsfield shared
the Nobel Prize in Medicine in 1979.
The four illustrations in Figs. 2.6 and 2.7 from the 1978 paper of Shepp and
Kruskal [588] show a mathematical phantom on the bottom in Fig. 2.6 representing
145
Fig. 2.6 On the bottom is a simulation of a human head using 11 ellipses. On the top is a
reconstruction using the algorithm embodied in the first commercial machine (EMI Ltd.) from
180 160 strip projection data obtained by exact calculation from the image on the bottom.
(From Shepp and Kruskal [588, pp. 422423]. Reprinted by permission of American Mathematics
Monthly)
146
Fig. 2.7 On the bottom is a reconstruction from the bottom image in Fig. 2.6 using the Fourier
based algorithm of Shepp. On the top is a reconstruction using the algorithm used in the 1970s
EMI machine from 180 239 strip projection data obtained by exact calculation from the bottom
image in Fig. 2.6. (From Shepp and Kruskal [588, p. 424]. Reprinted by permission of American
Mathematics Monthly
147
A description of the progress made between the 1970s and 1980s is contained in
the following quotation from Science 214 (1981), p. 1327:
When CT scanners first became commercially available about 8 years ago it took 5 minutes
to scan a patients head and 5 minutes for each computerized reconstruction of an image
from the x-ray data. Now, because of advances in the design of the scanners and in computer
technology, the newest machines can scan a head just 10 seconds and can reconstruct an
image virtually instantaneously. According to Jay Thomas Payne of Abbott Northwestern
Hospital in Minneapolis, the Mayo Clinics first CT scanner, which is only 5 years old, has
been relegated to the clinics historical museum.
Of course, despite the progress in speed and accuracy of CT scanners, they still
expose people to radiation. In a New York Times article from August 21, 2012, Jane
E. Brody wrote the following.
But it [radiation] also has a potentially serious medical downside: the ability to damage
DNA and, 10 to 20 years later, to cause cancer. CT scans alone, which deliver 100 to
500 times the radiation associated with an ordinary X-ray and now provide three-fourths of
Americans radiation exposure, are believed to account for 1.5 percent of all cancers that
occur in the United States.
Thus nuclear magnetic resonance tomography (NMR alias MRI) may be destined
to be the tomography of the future. It uses magnetic fields rather than x-rays and thus
is presumably less damaging to the body. Louis [431,432] considers applications of
the three-dimensional Radon transform to NMR tomography. Spherical harmonics
are used to improve the algorithm by studying the kernel (i.e., inverse image of
0) of the transform (ghosts). In 2003, the Nobel Prize for Physiology or Medicine
went to Paul Lauterbur and Sir Peter Mansfields for their work on MRI. See Elena
Prestini [522], Chap. 8, for more details on the subject. She begins the chapter
with a quotation from a New York Post article from 1939 about a talk by I. I. Rabi
explaining why NMR scanning works. The quote is: We are all radio stations.
There are many other applications of the Radon transform; for example, in radio
astronomy (see Bracewells article in Herman [290, pp. 81104]). And there are
applications to partial differential equations; e.g., to find solutions of the wave
equation (see Dym and McKean [147, pp. 137139], and Helgason [279].
This concludes our discussion of harmonic analysis on the sphere. It would be
possible to consider spherical analogues of many more results from Chap. 1. We
shall leave this to the interested reader. For example, the group of isometries of the
sphere is just O(3). A discontinuous subgroup O(3) has the property that any
domain in the sphere can contain only finitely many points equivalent under to
any given point. There are very few such discontinuous for the sphere or elliptic
plane. They correspond to the regular polyhedra (see Hilbert and Cohn-Vossen [297,
p. 242]). These are the analogues of the space groups considered in Sect. 1.4. It is
also possible to discuss the analogue of the Poisson summation formula for \S2
=
\G/K, G = SO(3), K = SO(2). We shall not do this here, since our main interest
is the Selberg trace formula for noncompact fundamental domains (see Chap. 3). In
148
fact, we never examined the Selberg trace formula when G is the Euclidean group
either (but see Hejhal [265]).
There is work on fast computation of Fourier transforms on the sphere by Driscoll
and Healy [140], for example. There are also finite analogues of Radon transforms
(see [668]).
http://www.springer.com/978-1-4614-7971-0