26.12.2012 Views

SOLVING INVERSE PROBLEMS FOR THE HAMMERSTEIN ...

SOLVING INVERSE PROBLEMS FOR THE HAMMERSTEIN ...

SOLVING INVERSE PROBLEMS FOR THE HAMMERSTEIN ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

International Journal of Pure and Applied Mathematics<br />

————————————————————————–<br />

Volume 60 No. 4 2010, 393-408<br />

<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong><br />

INTEGRAL EQUATION AND ITS RANDOM ANALOG<br />

USING <strong>THE</strong> “COLLAGE METHOD” <strong>FOR</strong> FIXED POINTS<br />

H.E. Kunze 1 § , D. La Torre 2 , K.M. Levere 3 , E.R. Vrscay 4<br />

1,3 Department of Mathematics and Statistics<br />

College of Physical and Engineering Science<br />

University of Guelph<br />

50, Stone Road East, Guelph, Ontario, N1G 2W1, CANADA<br />

1 e-mail: hkunze@uoguelph.ca<br />

3 e-mail: klevere@uoguelph.ca<br />

2 Department of Economics, Business and Statistics<br />

University of Milan<br />

Milan, ITALY<br />

e-mail: davide.latorre@unimi.it<br />

4 Department of Applied Mathematics<br />

University of Waterloo<br />

Waterloo, Ontario, N2L 3G1, CANADA<br />

e-mail: ervrscay@math.uwaterloo.ca<br />

Abstract: Many inverse problems in applied mathematics can be formulated<br />

as the approximation of a target element u in a complete metric space (X,dX)<br />

by the fixed point ¯x of an appropriate contraction mapping T : X ↦→ X. The<br />

method of collage coding seeks to solve this problem by finding a contraction<br />

mapping T that minimizes the so-called collage distance d(x,Tx). In this paper,<br />

we develop a collage coding framework for inverse problems involving deterministic<br />

or random Hammerstein integral operators. Such operators are used to<br />

model image blurring. We illustrate the method with examples.<br />

AMS Subject Classification: 45Q05, 45R05, 60H25<br />

Key Words: Hammerstein integral equations, inverse problems, fixed point<br />

equations, random fixed point equations, collage theorem<br />

Received: March 6, 2010 c○ 2010 Academic Publications<br />

§ Correspondence author


394 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

1. Introduction<br />

In this paper, we examine the Hammerstein integral equation in the context of<br />

Banach’s celebrated contraction mapping theorem. The Hammerstein integral<br />

operator is<br />

(Tu)(x) =<br />

� 1<br />

0<br />

K(x,y,u(y))dy + l(x).<br />

In imaging applications, u represents an input signal (or image) and the kernel<br />

K describes how this signal has been blurred. Of course, the kernel can take<br />

on many forms, as an image might be blurred for a wide variety of reasons<br />

ranging from effects of motion capture to focus problems to diffraction effects<br />

of the medium in which a photograph was taken. The term l(x) represents the<br />

contribution of additive noise to the recorded image. The result of applying the<br />

Hammerstein integral operator, Tu, is a blurred and noised image.<br />

Normally, Banach’s Theorem is used to establish the existence and uniqueness<br />

of solutions to the related fixed point equation u = Tu. Our objective,<br />

however, is to show that it can be used to solve the associated inverse problem:<br />

given a recorded signal as well as l, denoise and deblur the image by<br />

recovering the parameters in K (and possibly its functional form).<br />

We also show that the random Hammerstein integral and its related inverse<br />

problem can be treated with a random version of Banach’s Theorem.<br />

As we have shown in previous works by Forte and Vrscay [7], [8], La Torre<br />

and Iacus [9], [10], and Kunze et al [12], [13], [14], [16], [17], a number of inverse<br />

problems in applied mathematics can be formulated as the approximation of<br />

a target element x in a complete metric space (X,dX) by the fixed point ¯x of<br />

an appropriate contraction mapping T : X ↦→ X. In practice, one considers a<br />

family of such contraction mappings Tλ, λ ∈ R n , with respective fixed points<br />

¯xλ, and seeks to minimize the approximation error � x − ¯xλ �. However, it is<br />

generally impractical to search directly for contraction maps Tλ that minimize<br />

this approximation error. Instead, we resort to the following simple consequence<br />

of Banach’s Fixed Point Theorem, referred to as the collage theorem in fractal<br />

imaging literature (see Barsley et al [3]):<br />

dX(x, ¯x) ≤ 1<br />

1 − c dX(x,Tx), for all x ∈ X, (1)<br />

where c is the contraction factor of T. One now seeks a contraction mapping<br />

Tλ that minimizes the so-called collage error dX(x,Tx). In other words, we<br />

look for a contraction mapping T that sends the target x as close as possible<br />

to itself. This is the essence of collage coding that has provided the basis of


<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong>... 395<br />

most, if not all, fractal image coding and compression methods (see Fischer [6],<br />

Forte and Vrscay [7], [8], and Lu [18]). Indeed, many problems in the parameter<br />

estimation literature (see, for example, Anderson [1] and Milstein [19]) can be<br />

formulated in such a collage coding framework [17], [14], [13], [12].<br />

The structure of the paper is as follows. In Section 2, we state some relevant<br />

mathematical results regarding contraction mappings and their fixed points,<br />

and the inverse problem of approximation by such fixed points. In Sections<br />

3 and 4, we examine Hammerstein-type integral equations and also present a<br />

collage coding method to solve inverse problems. After discussing the deterministic<br />

case in these two sections, we present a parallel discussion in Sections<br />

4-6, ending in the latter section with examples of the solution of the inverse<br />

problem for the random Hammerstein integral equation via the collage theorem<br />

for random fixed point equations.<br />

2. Fixed Point Equations and the “Collage Theorem”<br />

For the benefit of the reader, we mention some important mathematical results.<br />

Definition 1. Let (X,dX) be a metric space. A mapping T : X ↦→ X is<br />

said to be Lipschitz on X if there exists a positive constant K > 0 such that<br />

dX(Tx,Ty) ≤ KdX(x,y), (2)<br />

for all x,y ∈ X. If the inequality is satisfied for K < 1, then T is said to be<br />

contractive. In this case we define its contraction factor c to be the infimum of<br />

all K values.<br />

Theorem 2.1. (Banach Fixed Point Theorem) Let (X,dX) be a complete<br />

metric space. Also let T : X ↦→ X be contractive. Then there exists a unique<br />

¯x ∈ X such that ¯x = T ¯x. Moreover, for any x ∈ X, dX(T n x, ¯x) ↦→ 0 as n ↦→ ∞.<br />

A simple triangle inequality argument along with Banach’s Theorem yields<br />

the following result.<br />

Theorem 2.2. (“Collage Theorem”, see Barnsley et al [2], [3]) Let (X,dX)<br />

be a complete metric space and T : X ↦→ X a contraction mapping with<br />

contraction factor c ∈ [0,1). Then for any x ∈ X,<br />

dX(x, ¯x) ≤ 1<br />

1 − c dX(x,Tx), (3)<br />

where ¯x is the fixed point of T.<br />

Theorem 2.3. (Continuity of Fixed Points, see Centore and Vrscay [5])


396 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

Let (X,dX) be a complete metric space and T1,T2 be two contractive mappings<br />

with fixed points ¯x1 and ¯x2. Then<br />

1<br />

dX(¯x1, ¯x2) ≤<br />

1 − min{c1,c2} dsup(T1,T2), (4)<br />

where<br />

dsup(T1,T2) = sup dX(T1(x),T2(x)) (5)<br />

x∈X<br />

and ci is the contractivity factor of Ti.<br />

Another manipulation of the triangle inequality involving x, Tx and ¯x yields<br />

the following interesting result.<br />

Theorem 2.4. (“Anti-Collage Theorem”, see Vrscay and Saupe [20])<br />

Assume the conditions of the collage theorem above. Then for any x ∈ X,<br />

dX(x, ¯x) ≥ 1<br />

1 + c dX(x,Tx). (6)<br />

3. The Hammerstein Integral Equation<br />

We are now interested in finding solutions of the fixed point equation u = Tu<br />

where T is a Hammerstein integral operator having the form<br />

(Tu)(x) =<br />

� 1<br />

0<br />

K(x,y,u(y))dy + l(x). (7)<br />

Here, l : [0,1] ↦→ R is a given function and K : R × R × R ↦→ R satisfies<br />

|K(α1,β,γ) − K(α2,β,γ)| ≤ C1(β,γ)|α1 − α2| (8)<br />

and |K(α,β,γ1) − K(α,β,γ2)| ≤ C2(α,β)|γ1 − γ2|. (9)<br />

Furthermore suppose that for u ∈ L 2 ([0,1]) the function ξ(y) := C1(y,u(y))<br />

belongs to L 2 ([0,1]). These next results are quite simple, and we mention them<br />

for the benefit of the reader. The first result states the existence and uniqueness<br />

of the solution in the space C([0,1]).<br />

Theorem 3.1. Let l : [0,1] ↦→ R be a continuous function, then:<br />

(i) (Tu)(x) is a continuous function.<br />

(ii) if �C2�∞ := sup α,β∈[0,1] |C2(α,β)| < 1 then T : C([0,1]) ↦→ C([0,1]) is a<br />

contraction.


<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong>... 397<br />

Proof. Given a continuous function u ∈ C([0,1]), we have<br />

|(Tu)(x) − (Tu)(z)| ≤<br />

� 1<br />

0<br />

≤ |x − z|<br />

≤ |x − z|<br />

|K(x,y,u(y)) − K(z,y,u(y))|dy + |l(x) − l(z)|<br />

� 1<br />

0<br />

�� 1<br />

0<br />

|C1(y,u(y))|dy + |l(x) − l(z)|<br />

|C1(y,u(y))| 2 �<br />

dy<br />

1<br />

2<br />

+ |l(x) − l(z)| (10)<br />

and this shows the continuity of the function (Tu)(x). To prove contractivity<br />

we have<br />

d∞(Tu,Tv) = sup<br />

≤ sup<br />

x∈[0,1]<br />

� 1<br />

0<br />

x∈[0,1]<br />

|(Tu)(x) − (Tv)(x)|<br />

|K(x,y,u(y))−K(x,y,v(y))|dy ≤ sup<br />

x∈[0,1]<br />

� 1<br />

Consider now the operator T on the space L 2 ([0,1]).<br />

0<br />

|C2(x,y)||u(y)−v(y)|dy<br />

≤ �C2�∞d∞(u,v). �<br />

Theorem 3.2. Suppose l ∈ C([0,1]) and C2 ∈ L 2 ([0,1] × [0,1]) with<br />

�C2�2 < 1. Then T : L 2 ([0,1]) ↦→ L 2 ([0,1]) is a contraction.<br />

Proof. We first prove that Tu ∈ L2 ([0,1]). For each fixed x1,x2 ∈ [0,1], we<br />

compute<br />

|Tu(x1) − Tu(x2)| ≤<br />

� 1<br />

0<br />

≤ |x1 − x2|<br />

≤ |x1 − x2|<br />

|K(x1,y,u(y)) − K(x2,y,u(y))|dy + |l(x1) − l(x2)|<br />

� 1<br />

0<br />

�� 1<br />

0<br />

|C1(y,u(y))|dy + |l(x1) − l(x2)|<br />

|C1(y,u(y))| 2 �<br />

dy<br />

1<br />

2<br />

+ |l(x1) − l(x2)|.<br />

So the function (Tu)(x) is bounded on [0,1] and then it belongs to L2 ([0,1]).<br />

To prove contractivity we have<br />

�Tu − Tv� 2 2 =<br />

≤<br />

� 1<br />

0<br />

�<br />

�<br />

�<br />

�<br />

� 1<br />

0<br />

≤<br />

� 1<br />

0<br />

|(Tu)(x) − (Tv)(x)| 2 dx<br />

K(x,y,u(y))dy + l(x) −<br />

� 1 �� 1<br />

0<br />

0<br />

� 1<br />

0<br />

�<br />

�<br />

K(x,y,v(y))dy − l(x) �<br />

�<br />

|K(x,y,u(y)) − K(x,y,v(y))|dy<br />

�2<br />

dx<br />

2<br />

dx


398 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

≤<br />

� 1 �� 1<br />

0<br />

0<br />

|C2(x,y)||u(y) − v(y)|dy<br />

≤<br />

� 1<br />

0<br />

�2<br />

dx<br />

�C2� 2 2�u − v�22 dx = �C2� 2 2�u − v�22 . �<br />

Corollary 3.1. If C2 ∈ L 2 ([0,1] × [0,1]) and �C2�2 < 1 then the fixed<br />

point equation Tu = u has a unique solution ū belonging to L 2 ([0,1]) and for<br />

all u0 ∈ L 2 ([0,1]) the sequence un = Tun−1 converges to ū. If �C2�∞ < 1 then<br />

ū is continuous.<br />

4. Inverse Problem for the Hammerstein Integral Equation<br />

For each fixed u ∈ R we have K(x,y, ·) ∈ L2 ([0,1] × [0,1]), and then<br />

∞�<br />

K(x,y,u) = ai,j(u)φj(x)φi(y), (11)<br />

i,j=1<br />

where φi(y) is an orthonormal basis in L 2 ([0,1]). For u ∈ C([0,1]), we suppose<br />

that the functions ai,j(u) satisfy:<br />

(i) ai,j(u) ∈ L2 ([umin,umax]) where umin = min<br />

0≤x≤1 u(x) and umax = max<br />

0≤x≤1 u(x),<br />

so that<br />

∞�<br />

ai,j(u) = ai,j,kψk(u),<br />

k=1<br />

where ψk(u) is a basis for L2 ([umin,umax]).<br />

�<br />

� ∞�<br />

�<br />

�<br />

�<br />

�<br />

(ii) Let |ai,j(u) − ai,j(v)| = � ai,j,k(ψk(u) − ψk(v)) �<br />

�<br />

�<br />

k=1<br />

≤ bi,j|u − v|. Then<br />

�<br />

�<br />

�<br />

�<br />

��<br />

�<br />

|K(x,y,u) − K(x,y,v)| = �<br />

� (ai,j(u) − ai,j(v))φj(x)φi(y) �<br />

�<br />

� i,j<br />

�<br />

�<br />

�<br />

�<br />

��<br />

�<br />

�<br />

≤ �<br />

� bi,jφj(x)φi(y) �<br />

� |u − v|,<br />

� i,j �<br />

�<br />

�<br />

�<br />

��<br />

�<br />

�<br />

so C2(x,y) = �<br />

� bi,jφj(x)φi(y) �<br />

� in (9).<br />

�<br />

�<br />

i,j<br />

We truncate each sum at upper limit n. Given a target solution u ∈ C([0,1])


<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong>... 399<br />

and the function l(x), the inverse problem consists of finding coefficients ai,j,k<br />

in the expansion of K so that u is approximately admitted as the fixed point<br />

of the corresponding Hammerstein integral operator. Our collage distance is<br />

�u − Tu� 2 2 =<br />

� � 1 � �<br />

� 1<br />

2<br />

�<br />

�<br />

�u(x) − K(x,y,u(y))dy − l(x) �<br />

� dx<br />

0 0<br />

�<br />

� � 1 �<br />

= �<br />

�<br />

0 � −<br />

�<br />

n�<br />

n�<br />

� 1 n�<br />

�2<br />

�<br />

ujφj(x) + ljφj(x) + ai,j,kψk(u(y))φi(y)φj(x)dy �<br />

� dx<br />

j=1 j=1<br />

0<br />

i,j,k=1<br />

�<br />

� ⎛<br />

⎞�<br />

� � 1 � n�<br />

n�<br />

� 1 n�<br />

�2<br />

�<br />

= �<br />

� φj(x) ⎝−uj + lj + ai,j,k ck,lφl(y)φi(y)dy ⎠�<br />

� dx<br />

0 �j=1<br />

i,k=1<br />

0<br />

l=1<br />

�<br />

� ⎛<br />

⎞�<br />

� � 1 � n�<br />

n�<br />

�2<br />

�<br />

= �<br />

� φj(x) ⎝−uj + lj + ai,j,kck,l⎠�<br />

� dx<br />

0 �j=1<br />

i,k,l=1 �<br />

⎡<br />

� 1 n�<br />

≤ ⎣ φ 2 ⎤ ⎡ ⎛<br />

⎞2⎤<br />

n�<br />

n�<br />

j(x) ⎦ ⎣ ⎝−uj + lj + ⎠ ⎦ dx<br />

0<br />

⎡<br />

= ⎣<br />

� 1<br />

0<br />

j=1<br />

j=1<br />

j=1<br />

n�<br />

φ 2 ⎤ ⎡ ⎛<br />

n�<br />

j(x)dx⎦<br />

⎣ ⎝−uj + lj +<br />

j=1<br />

= n<br />

i,k,l=1<br />

ai,j,kck,l<br />

n�<br />

i,k,l=1<br />

⎛<br />

n�<br />

⎝−uj + lj +<br />

j=1<br />

ai,j,kck,l<br />

n�<br />

i,k,l=1<br />

⎞2⎤<br />

⎠<br />

⎦<br />

ai,j,kck,l<br />

where ck,l = ck,l(u1,...,un) have known values since ψk are chosen and u = u(y)<br />

is our known target function. Regarding the constraints we have that:<br />

� 1 � 1<br />

(C2(x,y)) 2 ⎛<br />

� 1 � 1<br />

dxdy = ⎝ �<br />

⎞2<br />

bi,jφj(x)φi(y) ⎠ dxdy<br />

0<br />

0<br />

=<br />

0 0<br />

� 1 � 1<br />

0<br />

+2<br />

= �<br />

i,j<br />

0 i,j<br />

� 1 � 1<br />

0<br />

b 2 i,j<br />

i,j<br />

�<br />

(bi,jφj(x)φi(y)) 2 dxdy<br />

0<br />

�<br />

i,j,h,k<br />

� 1<br />

+ 2 �<br />

i,j,h,k<br />

φj(x)φi(y)φh(x)φk(y)dxdy<br />

0<br />

φj(x)φh(x)dx<br />

� 1<br />

0<br />

⎞<br />

⎠<br />

2<br />

φi(y)φk(y)dy<br />

,


400 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

= �<br />

b 2 i,j .<br />

So now the inverse problem becomes:<br />

i,j<br />

min �u − Tu�<br />

A 2 , (12)<br />

where A = { �<br />

i,j b2i,j ≤ CA < 1}.<br />

Example. We set l(x) = x and set Ktrue(x,y,u) = ktrue(x,y)u. In this<br />

case, ai,j(u) = ai,ju, so that<br />

∞�<br />

Ktrue(x,y,u) = ai,jφj(x)φi(y),<br />

�u − Tu� 2 2 ≤<br />

and bi,j = ai,j. We define<br />

Let<br />

⎧<br />

⎪⎨<br />

ktrue(x,y) =<br />

⎪⎩<br />

i,j=1<br />

n�<br />

�<br />

−uj + lj +<br />

i=1<br />

φ(·) ∈ {φs(·),φc(·)} =<br />

n�<br />

i=0<br />

ai,juj<br />

� 2<br />

π(−1 − x + y), x − y < − 1<br />

2 ,<br />

π(x − y), − 1 1<br />

≤ x − y ≤<br />

2 2 ,<br />

1<br />

π(1 − x + y), < x − y ≤,<br />

2<br />

�√ √ �<br />

2 sin(π·), 2 cos(π·) . (13)<br />

In this rather limited finite basis, the representation of ktrue(x,y) is<br />

kbasis(x,y) = 2<br />

π φs(x)φc(y) − 2<br />

π φc(x)φs(y) = 4<br />

sin(π(x − y)).<br />

π<br />

We build the Hammerstein integral operator<br />

(Ttrueu) (x) =<br />

� 1<br />

0<br />

ktrue(x,y)u(y)dy + l(x),<br />

and construct an approximation of its fixed point ūtrue(x) by iterating the map<br />

20 times on the initial function u0(x) = 0. We set utarget(x) = T 20u0(x). See<br />

Figure 1. Thus, our example inverse problem is: Given utarget(x) and l(x), find<br />

a representation of the kernel k(x,y) in our finite basis such that the corresponding<br />

Hammerstein integral operator admits utarget(x) as an approximate<br />

solution. Following our development, we minimize the constrained squared L2 collage distance<br />

∆ 2 = �u − Tu� 2 ⎛<br />

2 + λ⎝<br />

�<br />

⎞<br />

− CA⎠,<br />

i,j<br />

a 2 i,j<br />

,


<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong>... 401<br />

Figure 1: The graph of z = ktrue(x,y), z = kbasis(x,y), and y =<br />

utarget(x)<br />

where CA ∈ (0,1) is a chosen fixed value. We denote the kernel that minimizes<br />

∆ 2 by kcollage(x,y), and we observe that the desired result is kcollage(x,y) =<br />

kbasis(x,y). To achieve this result, we would have to set<br />

CA = 2<br />

� 2<br />

π<br />

� 2<br />

≈ 0.81057.<br />

Instead, we systematically set CA = i<br />

10 , i = 1,2,... ,9, and minimize each<br />

corresponding ∆2 . Perhaps, one might anticipate that the minimum of the<br />

nine minimal collage distances we find will occur when CA = 0.8.<br />

In Table 1, we present the results obtained when we restrict our basis fur-


402 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

ther, so that it consists of precisely the two functions ψ1(x,y) = 2sin(πx)cos(πy)<br />

and ψ2(x,y) = 2cos(πx)sin(πy), which appear in ktrue(x,y). The pleasing re-<br />

CA minimal ∆ 2 a1 a2<br />

0.1 0.17153 0.29898 −0.10301<br />

0.2 0.12759 0.40955 −0.17964<br />

0.3 0.09614 0.48283 −0.25860<br />

0.4 0.07136 0.53421 −0.33856<br />

0.5 0.05088 0.57113 −0.41690<br />

0.6 0.03356 0.59838 −0.49187<br />

0.7 0.01929 0.61908 −0.56280<br />

0.8 0.01120 0.63526 −0.62964<br />

0.9 0.01613 0.64823 −0.69267<br />

Table 1: Collage coding results with kcollage(x,y) = a1φ1(x,y) +<br />

a2φ2(x,y) and a2 1 + a22 = CA<br />

sult is that the minimum of the minimal collage distances occurs when CA = 0.8.<br />

Indeed,<br />

a1 and − a2 ≈ 2<br />

≈ 0.63661,<br />

π<br />

the true best result.<br />

Note that when we minimize ∆ 2 , treating CA as an additional variable, we<br />

obtain λ = 0, a linear system in the coefficients a1 and a2, and a constraint<br />

equation, as opposed to an inequality. Still, upon solving the linear system for<br />

a1 and a2, we can check whether these minimizing values lead to a value of<br />

CA < 1. For this example, we obtain a1 = 0.63644 and a2 = 0.63675, giving<br />

CA = 0.81051.<br />

When we add the basis function φs(2x)φs(y) with coefficient a3, we obtain<br />

a1 = 0.63675, a2 = −0.55321, and a3 = 0.04580, with CA = 0.71359. The<br />

quality of the approximation diminishes in part because in this example �C2�2 =<br />

0.90690, meaning the operator is not very contractive.<br />

If we modify κtrue, replacing each π by π<br />

8 , then the new value of �C2�2 is the<br />

smaller 0.113362. We obtain a1 = 0.07957, a2 = −0.07495, and a3 = 0.00722,<br />

compared to the desired values a1 = −a2 = 1<br />

4π ≈ 0.07958 and a3 = 0.


<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong>... 403<br />

5. Random Fixed Point Equations and the<br />

“Random Collage Theorem”<br />

Again let (X,dX) denote a complete metric space. If (X,dX) is separable,<br />

then it is said to be a Polish space. We also let (Ω, F,µ) be a probability<br />

space. A mapping T : Ω × X ↦→ X is called a random operator if for any<br />

x ∈ X the function T(·,x) is measurable. The random operator T is said to<br />

be continuous/Lipschitz/contractive if, for a.e. ω ∈ Ω, we have that T(ω, ·) is<br />

continuous/Lipschitz/contractive (see Bharucha and Reid [4]).<br />

A measurable mapping x : Ω ↦→ X is called a random fixed point of the<br />

random operator T if x is a solution of the equation<br />

T(ω,x(ω)) = x(ω), a.e. ω ∈ Ω. (14)<br />

Consider the space Y of all measurable functions x : Ω ↦→ X. If we define the operator<br />

˜ T : Y ↦→ Y as ( ˜ Ty)(ω) = T(ω,x(ω)), then the solutions of this fixed point<br />

equation on Y are the solutions of the random fixed point equation T(ω,x(ω)) =<br />

x(ω). Suppose that the metric dX is bounded, that is, dX(x1,x2) ≤ K for all<br />

x1,x2 ∈ X. Then the function ψ(ω) = dX(x1(ω),x2(ω)) : Ω ↦→ R is an element<br />

of L1 (Ω) for all x1,x2 ∈ Ω. Define the following function over the space Y × Y :<br />

�<br />

dY (x1,x2) = dX(x1(ω),x2(ω))dPω = E(dX(x1,x2)). (15)<br />

Ω<br />

In Kunze et al [15], we proved the following two results.<br />

Theorem 5.1. The space (Y,dY ) is a complete metric space.<br />

Theorem 5.2. Suppose that:<br />

(i) for all x ∈ Y the function ξ : Ω ↦→ X1 defined by ξ(ω) := T(ω,x(ω)),<br />

∀ ω ∈ Ω belongs to Y ,<br />

(ii) dY ( ˜ Tx1, ˜ Tx2) ≤ cdY (x1,x2) with c < 1.<br />

Then there exists a unique solution of ˜ T ¯x = ¯x, that is, T(ω, ¯x(ω)) = ¯x(ω)<br />

for a.e. ω ∈ Ω.<br />

Hypothesis (i) can be avoided if X is a Polish space, in which case the<br />

following result (see Itoh [11]) holds.<br />

Theorem 5.3. Let X be a Polish space, that is, a separable complete<br />

metric space, and T : Ω × X ↦→ X be a mapping such that for each ω ∈ Ω<br />

the function T(ω,.) is c(ω)-Lipschitz and for each x ∈ X the function T(.,x)<br />

is measurable. Let x : Ω ↦→ X be a measurable mapping; then the mapping<br />

ξ : Ω ↦→ X defined by ξ(ω) = T(ω,x(ω)) is measurable.


404 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

Corollary 5.1. Let T : Ω × X ↦→ X be a mapping such that for each<br />

ω ∈ Ω the function T(ω,.) is a c(ω)-contraction. Suppose that for each x ∈ X<br />

the function T(.,x) is measurable. Then there exists a unique solution of the<br />

equation ˜ T ¯x = ¯x that is T(ω, ¯x(ω)) = ¯x(ω) for a.e. ω ∈ Ω.<br />

As a consequence of the collage and continuity theorems, we have the following.<br />

Corollary 5.2. Suppose that the hypotheses of Theorem 5.2 are satisfied.<br />

Then for any x ∈ Y ,<br />

1<br />

1 + c dY (x, ˜ Tx) ≤ dY (x, ¯x) ≤ 1<br />

1 − c dY (x, ˜ Tx), (16)<br />

where ¯x is the fixed point of ˜ T, that is, ¯x(ω) := T(ω, ¯x(ω)), ω ∈ Ω.<br />

6. Inverse Problem for the Random Hammerstein Integral Equation<br />

We now consider the case when the kernel K and the function l also depend on<br />

ω ∈ Ω, where (Ω, F,p) is a probability space. Suppose that for a.e. ω ∈ Ω, we<br />

have<br />

P (w : |K(ω,α1,β,γ) − K(ω,α2,β,γ) ≤ C1(β,γ)|α1 − α2|) = 1, (17)<br />

and P (w : |K(ω,α,β,γ1) − K(ω,α,β,γ2) ≤ C2(α,β)|γ1 − γ2|) = 1, (18)<br />

and that for u(y) ∈ L2 ([0,1]) we have C1(u,u(y)) ∈ L2 ([0,1]). Furthermore,<br />

assume that l(ω, ·) ∈ C([0,1]) and that �C2�2 < 1 for a.e. ω ∈ Ω. Then for<br />

a.e. ω ∈ Ω, there is a unique function ū(ω,x) ∈ Ω × L2 ([0,1]) such that<br />

ū(ω,x) = T(ω,x,ū(ω,x)) =<br />

� 1<br />

0<br />

K(ω,x,y,ū(ω,y))dy + l(ω,x). (19)<br />

The inverse problem can be formulated as: Given a function u : Ω ↦→ L 2 ([0,1])<br />

and a family of operators ˜ Tλ : Y ↦→ Y , where Y = {all measurable functions x :<br />

Ω ↦→ X}, find λ such that u is the solution of random fixed point equation<br />

˜Tλu = u, (20)<br />

or, equivalently,<br />

Tλ(ω,ū(ω)) = ū(ω). (21)<br />

Example. Extending our deterministic example, we set l(x,ω) = a(ω)x,<br />

where a(ω) is a random variable with known mean µa and variance σ2 a , and<br />

define<br />

kbasis(x,y) = ksc(ω)φs(x)φc(y) − kcs(ω)φc(x)φs(y),


<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong>... 405<br />

where ksc(ω) and kcs(ω) are random variables with unknown means µsc and µcs<br />

and variances σ 2 sc and σ 2 cs, respectively. Given data from realizations u(ωj,x),<br />

j = 1,... ,N, of the random variable u(ω,x), the fixed point of the corresponding<br />

T, along with l(x,ω), we wish to recover the means and variances of the<br />

kernel parameters.<br />

Each of the realizations u(ωj,x), j = 1,... ,N, of the random variable<br />

u(ω,x) is the solution of a fixed point equation<br />

u(ωj,x) =<br />

� 1<br />

0<br />

(ksc(ωj)φs(x)φc(y) − kcs(ωj)φc(x)φs(y)) u(ωj,y)dy + a(ωj)x.<br />

Now, for each of our target functions, u(ωj,x), we can find the constant<br />

values ksc(ωj) and kcs(ωj) via the collage coding method for the deterministic<br />

Hammerstein integral equation, as in Section 4. Upon treating each of our<br />

realizations, we will have found values for ksc(ωj) and kcs(ωj) for j = 1,... ,N.<br />

We then construct the approximations<br />

N�<br />

N�<br />

µsc ≈ µsc,N = 1<br />

N<br />

j=1<br />

ksc(ωj) and µcs ≈ µcs,N = 1<br />

N<br />

j=1<br />

kcs(ωj). (22)<br />

Note that the results obtained from collage coding each realization are independent.<br />

Using these approximations of the means, we can also calculate the<br />

variances:<br />

σ 2 sc ≈ σ 2 sc,N =<br />

and σ 2 cs ≈ σ2 cs,N =<br />

1<br />

N − 1<br />

1<br />

N − 1<br />

N�<br />

(ksc(ωj) − µsc,N) 2<br />

j=1<br />

N�<br />

(kcs(ωj) − µcs,N) 2 .<br />

Some results are presented in Table 2. The first three columns present the<br />

distributions of a, ksc, and kcs used to generate the realizations. The final two<br />

columns present the approximations of the means and variances obtained by<br />

our collage method for the various values of N specified in the middle column.<br />

In all cases, we have treated the constraint value CA as a parameter when<br />

minimizing each collage objective function.<br />

Our method correctly identifies the distribution of the random variable and<br />

finds good approximations of the mean and variance. As expected, a larger<br />

sample size produces better estimates of the mean and variance.<br />

j=1


406 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

True Values Collage Method Approximations<br />

a ksc kcs N ksc kcs<br />

N(1, 0.01) N (0.6, 0.0004) N (0.6,0.0004) 10 (0.62681,0.00067) (-0.62649,0.00066)<br />

N(1, 0.01)<br />

exp(1)<br />

N (0.6, 0.0004)<br />

exp<br />

N (0.6,0.0004) 100 (0.63655,0.00065) (-0.63613,0.00063)<br />

` ´<br />

π<br />

exp ` ´<br />

π<br />

10 (0.46817,0.41222) (-0.46817,0.41222)<br />

2<br />

exp(1) exp ` ´ π<br />

2<br />

2<br />

exp ` ´ π<br />

2<br />

100 (0.65473,0.42717) (-0.64155,0.37227)<br />

Table 2: Results for the random Hammerstein integral operator inverse<br />

problem, approximations are in the form (mean, variance)<br />

Acknowledgments<br />

This work was developed during a research visit by D. La Torre to the Department<br />

of Applied Mathematics of the University of Waterloo. D. La Torre<br />

thanks ERV for this opportunity. For D. La Torre this work has been supported<br />

by COFIN Research Project 2004. This work has also been supported in part<br />

by research grants (H.E. Kunze and E.R. Vrscay) and a Canadian Graduate<br />

Scholarship (K.M. Levere) from the Natural Sciences and Engineering Research<br />

Council of Canada (NSERC), which are hereby gratefully acknowledged.<br />

References<br />

[1] D. Anderson, Compartmental Modeling and Tracer Kinetics, Springer,<br />

Berlin (1983).<br />

[2] M.F. Barnsley, Fractals Everywhere, Academic Press, New York (1989).<br />

[3] M.F. Barnsley, V. Ervin, D. Hardin, J. Lancaster, Solution of an inverse<br />

problem for fractals and other sets, Proc. Nat. Acad. Sci., USA, 83 (1985),<br />

1975-1977.<br />

[4] A.T. Bharucha-Reid, Random integral equations, Mathematics in Science<br />

and Engineering, Volume 96, Academic Press, New York-London (1972).<br />

[5] P. Centore, E.R. Vrscay, Continuity of fixed points for attractors and invariant<br />

meaures for iterated function systems, Canadian Math. Bull., 37<br />

(1994), 315-329.<br />

[6] Y. Fisher, Fractal Image Compression, Theory and Application, Springer-<br />

Verlag, NY (1995).


<strong>SOLVING</strong> <strong>INVERSE</strong> <strong>PROBLEMS</strong> <strong>FOR</strong> <strong>THE</strong> <strong>HAMMERSTEIN</strong>... 407<br />

[7] B. Forte, E.R. Vrscay, Theory of generalized fractal transforms, Fractal<br />

Image Encoding and Analysis, NATO ASI Series F, Volume 159 (Ed. Y.<br />

Fisher), Springer-Verlag, New York (1998).<br />

[8] B. Forte, E.R. Vrscay, Inverse problem methods for generalized fractal<br />

transforms, Fractal Image Encoding and Analysis, ibid.<br />

[9] S.M. Iacus, D. La Torre, Approximating distribution functions by iterated<br />

function systems, Journal of Applied Mathematics and Decision Sciences,<br />

1 (2005), 334-345.<br />

[10] S.M. Iacus, D. La Torre, A comparative simulation study on the IFS distribution<br />

function estimator, Nonlinear Analysis: Real World Applications,<br />

6 (2005), 774-785.<br />

[11] S. Itoh, Random fixed point theorems with an application to random differential<br />

equations in banach spaces, J. Math. Anal. App., 67 (1979), 261-273.<br />

[12] H. Kunze, D. Crabtree, Using Collage Coding to Solve Inverse Problems in<br />

Partial Differential Equations, Maplesoft Conference Proceedings (2005).<br />

[13] H. Kunze, S. Gomes, Solving an inverse problem for Urison-type integral<br />

equations using Banach’s Fixed Point Theorem, Inverse Problems, 19<br />

(2003), 411-418.<br />

[14] H. Kunze, J. Hicken, E.R. Vrscay, Inverse problems for ODEs using contraction<br />

maps: Suboptimality of the “Collage Method”, Inverse Problems,<br />

20 (2004), 977-991.<br />

[15] H. Kunze, D. La Torre, E.R. Vrscay, Random fixed point equations and<br />

inverse problems by collage method, UNIMI - Research Papers in Economics,<br />

Business and Statistics, www.bepress.edu/unimi.<br />

[16] H. Kunze, S. Murdock, Solving inverse two-point boundary value problems<br />

using collage coding, Inverse Problems, 22 (2006), 1179-1190.<br />

[17] H. Kunze, E.R. Vrscay, Solving inverse problems for ordinary differential<br />

equations using the Picard contraction mapping, Inverse Problems, 15<br />

(1999), 745-770.<br />

[18] N. Lu, Fractal Imaging, Academic Press, NY (2003).


408 H.E. Kunze, D. La Torre, K.M. Levere, E.R. Vrscay<br />

[19] J. Milstein, The inverse problem: Estimation of kinetic parameters, In:<br />

Modeling of Chemical Reaction Systems (Ed-s: K. Ebert, P. Deuflhard,<br />

W. Jäger), Springer, Berlin (1981).<br />

[20] E.R. Vrscay, D. Saupe, Can one break the “collage barrier” in fractal image<br />

coding?, In: Fractals: Theory and Applications in Engineering (Ed-s: M.<br />

Dekking, J. Levy-Vehel, E. Lutton, C. Tricot), Springer-Verlag, London<br />

(1999), 307-323.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!