Chapter 4

Analysis for Inverse Problem

Inverse problems are ill-posed when measurable data are either insufficient for uniqueness or insensitive to perturbations of parameters to be imaged. To solve an inverse problem in a robust way, we should adopt a reasonably well-posed modified model at the expense of a reduced spatial resolution and/or add additional a priori information. Finding a well-posed model subject to practical constraints of measurable quantities requires deep knowledge about various mathematical theories in partial differential equations (PDEs) and functional analysis, including uniqueness, regularity, stability, layer potential techniques, micro-local analysis, regularization, spectral theory and others. In this chapter, we present various mathematical techniques that are frequently used for rigorous analysis and investigation of quantitative properties in forward and inverse problems.

4.1 Examples of Inverse Problems in Medical Imaging

Most inverse problems in imaging are to reconstruct cross-sectional images of a material property P from knowledge of input data X and output data Y. We express its forward problem in an abstract form as

4.1 4.1

where F is a nonlinear or linear function of P and X. To treat the problem in a computationally manageable way, we need to figure out its sensitivity, explaining how a perturbation P + ΔP influences the output data Y + ΔY. In this section, we briefly introduce some examples of forward problem formulations in imaging electrical and mechanical material properties of an imaging object such as the human body.

4.1.1 Electrical Property Imaging

We assume that an imaging object occupies a domain Ω in the three-dimensional space images. When the object is a human body, we are interested in visualizing electrical properties of biological tissues inside the body. We denote the admittivity of a biological tissue at a position r and angular frequency ω as images, where σω(r) and images are the conductivity and permittivity, respectively. Most biological tissues are resistive at low frequencies but capacitive terms may not be negligible at 10 kHz or above. Electrical property imaging aims to image the internal admittivity distribution P = γω(r), which depends on r = (x, y, z) and ω. To measure the passive material property, we must employ a probing method, which excites the object by externally applying a form of energy and measures its response affected by the admittivity.

As a probing method, we inject a sinusoidal current of Isin(ωt) mA through a pair of electrodes that are attached on the surface Ω. This sinusoidally time-varying current produces sinusoidal variations of electric and magnetic fields at every point r with the same angular frequency ω. For these sinusoidal fields, it is convenient to use the phasor notation. We denote the sinusoidally time-varying electric and magnetic fields by images and images, respectively. We express them using vector field phasors of E(r) and H(r) as

images

Each component of E(r) or H(r) is a complex-valued function (independent of time) that contains amplitude and phase information. Table 4.1 summarizes important variables and parameters in the steady-state or time-harmonic electromagnetic field analysis.

Table 4.1 Time-varying and time-harmonic electromagnetic fields in Maxwell's equations

Name Time-varying field Time-harmonic field
Gauss's law images images
Gauss's law for magnetism images ∇ · B = 0
Faraday's law of induction images ∇ × E = − iωB
Ampère's circuit law images ∇ × H = σE + iωD

For the admittivity imaging, it is imperative to produce a current density J inside Ω to sense the admittivity by Ohm's law images, where E is the electric field intensity. To create J(r) and E(r) inside the body Ω, we can use a pair of electrodes (attached on the surface of the body) to inject sinusoidal current. This produces J, E, B inside Ω, where B is the magnetic flux density. Alternatively, we may use an external coil to produce eddy currents inside the body Ω. In both cases, we should measure some quantities that enable us to estimate J and E (directly or iteratively). Here, we use Maxwell's equations to establish relations among J, E, input data and measured data.

We inject current of Isin(ωt) mA with 0 ≤ ω/2π ≤ 100 kHz through a pair of surface electrodes images and images on ∂Ω. The diameter of the imaging object is less than 1 m. Since the Faraday induction is negligibly small in this case, E, J and H approximately satisfy

4.2 4.2

4.3 4.3

with the boundary conditions

4.4 4.4

4.5 4.5

4.6 4.6

where n is the outward unit normal vector on ∂Ω and dS is the area element on ∂Ω. The first boundary condition (4.4) comes from J = γωE = 0 in the air, where γω ≈ 0. The boundary condition (4.5) means that the vector J on each electrode is parallel or antiparallel to n since the electrodes are highly conductive. One may adopt the Robin boundary condition to include effects of the electrode–skin contact impedance.

Since ∇ × E ≈ 0, there exists a scalar potential u such that

images

Since 0 = ∇ · (∇ × H) = ∇ · J = − ∇ · (γωu), the potential u is a solution of the following boundary value problem:

4.7 4.7


Example 4.1.1
Electrical impedance tomography (EIT) aims to visualize (the time or frequency change of) the admittivity distribution P = γ. One may use multiple electrodes images attached on the boundary ∂Ω to inject currents and to measure boundary voltages. The input X could be a sequence of injection currents through chosen pairs of electrodes, and the output Y is the resulting boundary voltage data set uj[σ]|∂Ω, j = 1, …, N − 1, where the potential uj[σ] is a solution of the Neumann boundary value problem (4.7) with images and images. If γ = γ0 + Δγ (or P = P0 + ΔP) is a small perturbation of a known quantity γ0, we may deal with a linearized problem:

images


 


Example 4.1.2
Magnetic resonance current density imaging (MRCDI) is an imaging technique that visualizes a current density vector distribution by using a magnetic resonance imaging (MRI) scanner. Measurements are performed by applying an external current to the imaging object during an MRI acquisition. In MRCDI, the quantity to be imaged is the internal current density P = J; the input X is an injection current through a pair of electrodes images and images; the output Y is the internal data of b that is measured by the MRI scanner. Note that measuring all three components of B = (Bx, By, Bz) requires rotating the imaging object inside the MRI scanner, since one can measure only one component of b, which is in the direction of the main magnetic field of the MR scanner. MRCDI uses the relation

images


 


Example 4.1.3
In magnetic resonance electrical impedance tomography (MREIT), the property to be imaged is the conductivity distribution P = σ at a low frequency; the input X is an injection current through a pair of electrodes images and images; the output Y is the internal data Hz(r) that is measured by an MRI scanner with its main field in the z direction. The relation between Hz and σ is determined by the z component of the Biot–Savart law:

4.8 4.81

Here, images is a harmonic function determined by the geometry of lead wire and electrodes and u is the voltage satisfying the Neumann boundary value problem (4.7) with γ replaced by σ.

 


Example 4.1.4
Magnetic resonance electrical property tomography (MREPT) aims to visualize P = γω at the Larmor frequency. In MREPT, the input X is a radio-frequency (RF) excitation by an external coil; the output Y is the positive rotating magnetic field images that is measured by a B1-mapping technique (Stollberger and Wach 1996; Akoka et al. 2009). The relation between γω and H+ is determined by

4.9 4.82


 


Example 4.1.5
Magnetic induction tomography (MIT) aims to provide the admittivity distribution P = γω in the frequency range of 1–10 MHz. In MIT, multiple transmit coils are used to excite the imaging object and the same or different receive coils are used to measure induced voltages by time-varying magnetic fields produced by eddy currents. The sinusoidal frequency of the external excitation must be high enough to produce a measurable induced voltage:

4.10 4.83

where C is the closed path of the receive coil and S is the coil surface. We note that the magnetic field H outside Ω conveys information on the admittivity γω inside the imaging object Ω. In MIT, the input X is the external excitation; the output Y is the measured induced voltage.

4.1.2 Mechanical Property Imaging

Elasticity is a mechanical property of an elastic object describing how the deformed object, subject to an external force, returns to its original state after the force is removed. Elastography measures the propagation of transverse strain waves in an object. Tissue stiffness is closely related to the velocity of the wave, and the shear modulus (or modulus of rigidity) varies over a wide range, differentiating various pathological states of tissues. Hence, the speed of the harmonic elastic wave provides quantitative information for describing malignant tissues, which typically are known to be much stiffer than normal tissues.

In imaging a mechanical property, we should use Hooke's law, which links the stress and strain tensors:

images

where (Cijkl) is the stiffness tensor, (σij) is the strain tensor and images is the stress tensor. We need to apply a mechanical vibration or stress to create a displacement in Ω and measure some quantity that enables us to estimate the displacement inside Ω. Then, we use an elasticity equation to connect the stress tensor, strain tensor, displacement, input data and measured data.

Assuming that we are trying to image a shear modulus distribution in a linearly elastic and isotropic material, the time-harmonic elastic displacement field denoted by u = (u1, u2, u3) is dictated by the following PDE:

4.11 4.8

where we use the following notation: ρ is the density of the material,

images

is the shear modulus,

images

is the Lamé coefficient, E is Young's modulus and σ is Poisson's ratio.


Example 4.1.6
Elastography has been used as a non-invasive technique for the evaluation of the stiffness of the liver, for example. The mechanical property to be imaged is the shear modulus P = μ; the input X is an external low-frequency mechanical vibration applied to the boundary of the imaging object; the output Y is an internal measurement of the displacement vector u. The relation between P = μ and Y = u is determined by the PDE (4.11). Magnetic resonance elastography (MRE) uses an MR scanner to measure the interior displacement, whereas transient elastography (TE) uses an ultrasound system.

4.1.3 Image Restoration

The image restoration problem is to recover an original image P = u from a degraded measured (or observed) image Y = f that are related by

4.12 4.9

where η represents noise and H is a linear operator including blurring and shifting. One may often use a priori knowledge about the structure of the true image that can be investigated by looking at geometric structure by level curves.

A typical way of denoising (or image restoration) is to find the best function by minimizing the functional

4.13 4.10

where the first fidelity term images forces the residual Huf to be small, the second regularization term images enforces the regularity of u and the regularization parameter λ controls the tradeoff between the residual norm and the regularity.

4.2 Basic Analysis

To deal with an imaging object with an inhomogeneous material property, we consider the following PDE:

4.14 4.11

where σ is differentiable and u is twice differentiable in Ω for ∇ · (σ(r)∇u(r)) = 0 to make sense with the classical derivatives.

In practice, the material property σ may change abruptly. For example, we may consider a conductivity distribution σ inside the human body Ω. Then, σ may have a jump along the boundary of two different organs. Along such a boundary, the electrical field E = − ∇u, induced by an injection current through a pair of surface electrodes attached on ∂Ω, may not be continuous due to interface conditions of the electric field (like the refractive condition of Snell's law). In this case, there exists no solution uC2(Ω) in the classical sense and we should seek a practically meaningful solution uC2(Ω). Hence, we need to expand the admissible set of solutions of (4.14), which must include all practically meaningful solutions.

This motivates the concept of the generalized derivative called the weak derivative, which is a natural extension of the classical derivative. With the use of the weak derivative (reflecting the refractive condition of Snell's law), we can manage the equation ∇ · (σ∇u) = 0 for a discontinuous σ. For a quick understanding of the weaker derivative, we consider the following one-dimensional Dirichlet problem:

4.15 4.12

Hence, u satisfies

images

We should note that u is different from the solution v of

images

This is because v is linear whereas the potential u is piecewise-linear satisfying the following transmission condition (the refractive condition) at x = 0 where σ is discontinuous:

4.16 4.13

Indeed, the practical solution of (4.15) is

images

Note that the classical derivative u′ does not exist at x = 0:

images

The difficulty regarding the refraction contained in the PDE

images

can be removed by the use of the variational framework:

4.17 4.14


Exercise 4.2.1
Show that a solution u of (4.17) satisfies the transmission condition (4.16).

We need to take account of the set of physically meaningful solutions of the variational problem (4.17). A practically meaningful solution u, which is a voltage, should have a finite energy, that is,

images

We solve (4.15) by finding u in the admissible set images Indeed, the solution u of (4.15) is a minimizer of Φ(v) within the set images:

4.18 4.15

This will be explained in section 4.3.

We return to the three-dimensional problem of (4.14), where u(r) represents an electrical potential at r in an electrically conducting domain Ω. The physically meaningful solution u must have a finite energy:

images

Hence, the solution of (4.14) should be contained in the set images. Assuming images, the set images is the same as the following Hilbert space:

images

where

images

Indeed, the solution is the minimizer of Φ(u) within the set {vH1(Ω):v|∂Ω = f}. As we mentioned before, the PDE − ∇ · (σ(r)∇u(r)) = 0 should be understood in the variational framework as

4.19 4.16


Remark 4.2.2
When images, there is no difference between the classical and variational problems. However, there exist reasonable situations where the variational problem lacks a smooth solution; there are many practical σ for which the minimization problem has no solution in the class images. Indeed, we can consider a problem with images. Obviously, the classical problem does not have a solution. We can construct a minimizing sequence {un} in images that is a Cauchy sequence with respect to the norm images. Although the Cauchy sequence {un} does not converge within images, it converges in the Sobolev space H1(Ω), the completion of images with respect to the norm images. This means that we can solve the minimization and variational problem within the Sobolev space H1(Ω).

4.2.1 Sobolev Space

To explain the solution of a PDE in the variational framework, it is convenient to introduce the Sobolev space. Let Ω be a bounded smooth domain in images with its smooth boundary ∂Ω and let x = (x1, …, xn) represent a position. We introduce a Sobolev space H1(Ω), which is the closure of the set images equipped with the norm

images

The finite element method (FEM) for computing an approximate numerical solution and its error analysis can be easily accomplished within the variational framework with the Hilbert space. This Hilbert space H1(Ω) is the most widely used Sobolev space in PDE theory.

The generalized derivative can be explained by means of the integration-by-parts formula:

images

In general,

images

where the notions ∂α and |α| are understood in the following way:

  • images, images and images;
  • images, for example, images;
  • |α| = ∑k=1nαk.

A function vi satisfying the following equality behaves like the classical derivative images:

images


Definition 4.2.3
Let images and 0 ≤ α ≤ 1. An open set images is said to be a Ck, α domain if, for each p ∈ ∂Ω, there exists an open neighborhood Up of p and a Ck, α diffeomorphism Φp:UpB1(0) such that

images


We will assume that Ω is an open subdomain of images with its Ck, α boundary ∂Ω, where k + α ≥ 1. We are now ready to introduce the Sobolev spaces Wk, p(Ω) and images, where Ω is a domain in images with its boundary ∂Ω and images:

  • W1, p(Ω) and images are the completion (or closure) of images and images, respectively, with respect to the norm

images

In other words,

images

  • W2, p(Ω) and images are the completion (or closure) of images and images, respectively, with respect to the norm

images

  • We denote images and images.

Example 4.2.4
Let Ω = (0, 1) and u(x) = x(1 − x) in Ω. We can show that images but images. To see this, assume that there exists images such that umu in H2(Ω) and images without loss of generality. For 0 < s < 1, it follows from the Schwarz inequality that

images

Hence,

images

An elementary computation shows that

images

Combining the above two inequalities leads to

4.20 4.84

In particular, substituting images into (4.20) yields

images

and therefore

images

This means that images, which contradicts the assumption. For details on this example, see the lecture note by Feldmann and Uhlmann (2003).

 


Example 4.2.5
There exists a function images that is not bounded on every non-empty open set in images. Denoting the set of rational numbers by images, there exists a sequence images. For

images

we can show by a direct computation that images but it is unbounded on every non-empty open set.

 


Example 4.2.6
Let images. Let images and images with images and tanθ = y/x. Show that u1 and u2 satisfy

images

Hence, w = u1u2 satisfies

images

According to the maximum principle, we obtain

images

But, we know that images. What is wrong with this conclusion? We should note that u1H1(Ω) and hence u1 is not a practically meaningful solution. Because of wH1(Ω), we cannot apply the maximum principle.

4.2.2 Some Important Estimates

We begin by explaining simplified versions of two important inequalities, the Poincaré and trace inequalities.


Example 4.2.7
Let Ω = {(x, y):0 < x, y < a} be a square with side length a > 0.
  • A simplified version of the Poincaré inequality is

images

  • A simplified version of the trace inequality is

images

These inequalities are based on the fundamental theorem of calculus:

images

From the first identity in the above expressions, we have

images

and therefore

images

Integrating over the y variable gives

images

and therefore

images

Application of the above estimate to the three other sides of ∂Ω leads to the trace inequality.
The Poincaré inequality uses the special property that u|∂Ω = 0 to get

images

and hence

images

which gives the Poincaré inequality.

 


Theorem 4.2.8 (Hardy–Littlewood–Sobolev inequality)
Let p, r > 1 and 0 < λ < n, with

images

For images and images,

images

where

images

λ* = λ/n and Sn−1 denotes the unit sphere.

For the proof, see Theorem 4.3 in the book by Lieb and Loss (2001).


Theorem 4.2.9 (Sobolev's inequality)
Let images. Then the following inequalities hold.
  • For n ≥ 3,

4.21 4.17

where

images

  • If images (one dimension), then

4.22 4.18

  • If images (two dimensions), then, for each images,

4.23 4.19

where

images

  • Let Ω be a C0, 1 domain and 1 ≤ pq, m ≥ 1 and km. Then,

4.24 4.20

4.25 4.21

where C is independent of u.

For the proof, see Theorem 8.8 in the book by Lieb and Loss (2001).


Lemma 4.2.10
Let Ω be a convex open set in images with diameter d and let E be an open subset of Ω. For images, we have

4.26 4.22

4.27 4.23

where uE = (1/|E|)∫Eu is the average of u over E and

images


 


Proof.
Let x ∈ Ω be fixed. We have

images

where θy = (yx)/|yx|. Hence,

images

Using the change of the order of integration and y = x + ρθ, we obtain

images

where images. The second inequality (4.27) follows from the Hardy–Littlewood–Sobolev inequality.

 


Definition 4.2.11
For images, define

images


This definition is consistent with the previous definition of Hk(Ω) for images. For the example of images, it follows from the Plancherel theorem that

images


Theorem 4.2.12
For given images with s ≥ 0, there exists images such that

images

Moreover, images and

images


 


Proof.
By the Riesz representation theorem, there exists images such that

images

where 〈 ·, · 〉 s is the inner product of Hs. Defining v by images we have images.

 


Exercise 4.2.13
Prove that, for each images, there exists images such that

images

where C1 and C2 are positive constants and independent of u. Note that we need to choose images such that

images

There are many choice of {f0, f1, f2}. For example, we could choose

images


 


Exercise 4.2.14
Let Ω be a domain in images. Prove that images if and only if images. This means that if images, then images.

 


Definition 4.2.15 (Negative Sobolev space)
The negative Sobolev space is images for s > 0 where images so that images.

 


Theorem 4.2.16
For each images, there exists {f0, f1, …, fn} ⊂ L2(Ω) such that

4.28 4.24

Moreover, images obeys (4.28).

 


Proof.
Let images be a bounded linear map defined by images. Since images, the map images is unitary. Hence, we can define a bounded linear map images by

images

From the Riesz representation theorem, there exists (f0, f1, …, fn) ∈ W such that

images

and images. For the complete proof, see the lecture note by Feldmann and Uhlmann (2003).

 


Remark 4.2.17
For vL2(Ω), the map images is defined by imagesv(u) = ∫Ωvu dx. Since images, we have images and images. Moreover, we can show that L2(Ω) is dense in H−1(Ω). By the Riesz representation theorem, for all imagesH−1(Ω), there exists images such that images for all images and images. Consider the map

images

If L2(Ω) is not dense in H−1(Ω), there exists non-zero images such that images for all vL2(Ω). Since images for all vL2(Ω), then w0 = 0. Hence, H−1(Ω) can be viewed as a completion of L2(Ω) with the norm

images


 


Remark 4.2.18
According to the Hahn–Banach theorem in section 4.4.2, we can extend images to images. But images is more useful than [H1(Ω)]* since (4.28) can be written as

images

provided that f0, …, fn are differentiable.

 


Definition 4.2.19
Let images be a images bounded domain. Then, there exist p1, …, pN ∈ ∂Ω, an open neighborhood Up of pj and a images diffeomorphism images such that images and

images

Choosing images such that images in a neighborhood of ∂Ω, we define Hs(∂Ω) by the closure of images with respect to the norm

images


 


Lemma 4.2.20 (Trace: restriction and extension)
Let images. There is a positive constant C depending only on s such that

images

On the other hand, for each images, there exists images such that

images


 


Proof.
From the uniform bounded principle theorem, it suffices to prove that

images

For simplicity of notation, we shall prove only the two-dimensional case n = 2. Indeed, we can apply the same idea to prove it for the general case. Let images be the projection map to the first component. From the Fourier inversion formula, we have

images

and

images

Comparing the above two identities, we have images and it follows from the Schwarz inequality that

images

Hence, we have

images

For the remaining proof, see the lecture note by Feldmann and Uhlmann (2003).

4.2.3 Helmholtz Decomposition

The Helmholtz decomposition states that any smooth vector field F in a smooth bounded domain Ω can be resolved into the sum of a divergence-free (solenoidal) vector field and a curl-free (irrotational) vector field.


Theorem 4.2.21
Every vector field F(r) = (F1(r), F2(r), F3(r)) ∈ [L2(Ω)]3 can be decomposed into

4.29 4.25

where u is a scalar function, ∇ · A = 0 and harmonic is a vector field whose Laplacian is zero in Ω. Moreover, u and A are solutions of2u = ∇ · F and2A = ∇ × F, with appropriate boundary conditions. Hence, these can be uniquely determined up to harmonic functions:

4.30 4.26

and

4.31 4.27


 


Proof.
We write the vector field F as

images

Integration by parts yields

4.32 4.28

where

images

Owing to the property

images

we have

images

Using the vector identity − ∇2F = ∇ × (∇ × F) − ∇(∇ · F), we can express (4.32) as

4.33 4.29

Integrating by parts again, we have

images

where Ψ2 is a function satisfying images for r ∈ Ω and r′ ∈ ∂Ω. This completes the proof of the Helmholtz decomposition.

4.3 Variational Problems

4.3.1 Lax–Milgram Theorem

Let Ω be a bounded domain in images with smooth boundary ∂Ω. In this section, we discuss the existence, uniqueness and stability of the boundary value problem (BVP):

4.34 4.30

where c0 and c1 are positive constants. In the case when σ is discontinuous, the fundamental mathematical theories of solutions such as existence, uniqueness and stability can be established within the framework of the Hilbert space H1(Ω), which is a norm closure of images with the norm images.

This problem (4.34) is equivalent to the following variational problem:

images

Define the map images by

4.35 4.31

and define images by

4.36 4.32

Hence, the solvability problem of (4.34) is equivalent to the uniqueness and existence question of finding images satisfying

4.37 4.33

Note that the map images satisfies the following:

  • both images and images are linear for all u, wX;
  • |a(u, v)| < c1||u|| ||v|| and c0||u||2a(u, u).

Here, the norm ||u|| is images. The estimate |a(u, v)| < c1||u|| ||v|| can be obtained using the Schwarz inequality and the Poincaré inequality gives the estimate c0||u||2a(u, u).

Assuming that the Hilbert space images has a basis {ϕk:k = 1, 2, …}, the variational problem (4.37) is to determine the coefficient {uk:k = 1, 2, …} of u = ∑ukϕk satisfying

4.38 4.34

Taking advantage of the linearity of a( ·, · ) and b( · ), the problem (4.38) is equivalent to solving

images

where A can be viewed as an images matrix with coefficients akj = ak, ϕj). From the fact that images, A is positive definite and invertible. Therefore, for given data b, there exists a unique u satisfying Au = b.

Now, we will prove the uniqueness and existence in images in a general setting. Let X be a Hilbert space over images with a norm || · ||. The map images is called a bounded bilinear map on the Hilbert space X if

  • both images and images are linear for all u, wX, and
  • there exist M so that |a(u, v)| < M||u|| ||v||.

The solvability of the problem (4.34) can be obtained from the following theorem.


Theorem 4.3.1 (Lax–Milgram theorem)
Suppose the bounded bilinear map images is symmetric and

4.39 4.35

Suppose images is linear.
1. There exists a unique solution uX of the following minimization problem:

4.40 4.36

2. The solution u of the minimization problem (4.40) is the solution of the following variational problem:

4.41 4.37


Before proving Theorem 4.3.1, we gain its key idea by simple examples.


Example 4.3.2
Let images, bX and A be an n × n symmetric matrix having positive eigenvalues. Define

images

Then, a( ·, · ) is bounded bilinear and b (·) is bounded linear. Since A has positive eigenvalues, a( ·, · ) satisfies the condition (4.39). Hence, it meets the requirements of Theorem 4.3.1. Noting that

images

we have

images


 


Example 4.3.3 (Laplace equation)
Let Ω be a bounded domain in images with a smooth boundary. We consider the following minimization problem:

images

This minimization problem has a unique solution in the Sobolev space images.
  • images is a Hilbert space with the norm images
  • images is bounded bilinear on X.
  • images is bounded linear on X when fL2(Ω).
  • a( ·, · ) is coercive (which will be proved in the next section):

images

  • From Theorem 4.3.1, there exists a unique solution images of the above minimization problem.
  • This minimizer u satisfies the variational problem:

images

  • This solution u is a solution of the corresponding Euler–Lagrange equation:

4.42 4.85

4.43 4.86


 


Proof.
(Lax–Milgram theorem) We begin by proving the existence of the minimization problem. Denoting

images

there exists a minimizing sequence {un}, that is,

images

Existence of the minimizer can be proved by showing that

4.44 4.38

The proof of (4.44) is based on the identity

4.45 4.39

This leads to

4.46 4.40

Setting images, we have

images

Hence {un} is a Cauchy sequence and there exists uX so that images. Since Φ( · ) is continuous, owing to the assumption that a( ·, · ) and b( · ) are bounded and linear, we prove the existence of the minimizer u:

images

Now, we prove uniqueness. If u is a minimizer,

images

Hence, if u and v are minimizers, then

images

which is equivalent to

images

In particular, images. Then, it follows from (4.39) that

images

which gives u = v.

 


Exercise 4.3.4
Prove that images.

4.3.2 Ritz Approach

Taking account of the finite element method, we continue to study Hilbert space techniques. As before, we assume the following:

  • X is a real Hilbert space with norm || · ||;
  • {Xn} is a sequence of a finite-dimensional subspaces of X such that XnXn+1 and

images

  • images is a basis of Xn;
  • images is a bounded, symmetric, strongly positive, bilinear map and

images

  • bX* where X* is the set of linear functionals on X.

As before, we consider the minimization problem:

4.47 4.41

The variational problem is as follows:

4.48 4.42

According to the Lax–Milgram theorem, we know the solvability of (4.47) and (4.48). We try to give a rather different analysis by using the new inner product a(u, v). Since ||u||2a(u, u), images can be viewed as a norm of the Hilbert space X.


Theorem 4.3.5 (Projection theorem)
The space X equipped with the new inner product and norm

4.49 4.43

is also a Hilbert space. For any closed subspace V = imagesX and for each uX,

4.50 4.44

Denoting images, if u is decomposed as

images

then u1 = uV and u2 = uuV.

 


Proof.
Let u be fixed. Since

images

the problem (4.50) is equivalent to solving the following minimization problem:

4.51 4.45

where b(v) = a(u, v). According to the Lax–Milgram theorem, there exists a unique solution uVV of the minimization problem (4.51). Next, we will show uuVV. Since

images

we have

images

Hence,

images

Letting t → 0, we get

images

and uuVV. Uniqueness of the decomposition follows from the fact that

images


 


Definition 4.3.6
Let X be a Hilbert space over the scalar field images. The dual of X, denoted by X*, is the set of linear maps images.

 


Theorem 4.3.7 (Riesz representation theorem)
For each linear functional bX*, there is a unique u*X such that

4.52 4.46

Moreover, images.

 


Proof.
Assume b ≠ 0, that is, b(v) ≠ 0 for some vX. We first prove that the null space V: = {vX:b(v) = 0} is a closed subspace of X of codimension 1, that is,

images

From the assumption of b ≠ 0 and using the linearity of b( · ),

images

Owing to the linearity of b( · ) and b(u) = 1, we have

images

Hence,

images

This eventually gives a decomposition:

images

According to the projection theorem, the above decomposition is unique, and this proves that dimV = 1. Setting u* = u/a(u, u), we have

images

From the above identity, it is easy to see that u* is unique. Since |b(v)| ≤ ||u*||a||v||a and images, we have ||b|| = ||u*||a. This completes the proof.

 


Theorem 4.3.8
Assume that unXn is a unique solution of the minimization problem (4.47) or the variational problem (4.48). Let u be a solution of the variational problem (4.48) with Xn replaced by X. Then,
i. images
ii. images
iii. (c/2)||uun||2 ≤ Φ(un) − Φ(u).

 


Proof.
Since a(u, v) = b(v) = a(un, v) for all vXn, we have

images

Since a(uun, uun) = a(uun, u) and a(uun, uv) = a(uun, u) for all vXn, we have

images

which gives (ii). Part (i) follows from the assumption that XnX as images. For all vX,

images

and therefore

images

which proves (iii).

 


Exercise 4.3.9 (Contraction mapping)
Let X be a Hilbert space over images. Assume an operator images is a contraction mapping, that is, there exists 0 < θ < 1 such that images for all u, vX. Then, there exists a unique u*X such that

images

Note that, for fixed u0X, we can define a sequence images and prove that the sequence is a Cauchy sequence.

 


Exercise 4.3.10 (Lax–Milgram theorem for non-symmetric bilinear operator)
Let X be a Hilbert space over images. Assume that the map images is bounded, coercive and bilinear:

images

Show that, for each bX*, there is a unique uX such that

images

Note that, for each vX, we can define a linear functional imagesvX* by imagesv(ϕ) = a(ϕ, v). We need to prove that there exists a unique uX such that imagesu = b. By the Riesz representation theorem, there exists a bounded linear operator images such that images for ϕ ∈ X and imagesX*. We need to show that there exists a unique u such that images. Define a map images by

images

For sufficiently small β > 0, images is a contraction map because

images

With 0 < β < c/(M2 + 1), there is a unique uX such that images, which is equivalent to images.

 


Exercise 4.3.11 (Lax–Milgram theorem on Hilbert space over images)
Let X be a Hilbert space over the complex scalar field images. Assume that the map images satisfies the following:
  • images is linear for each uX;
  • images for all u, wX;
  • there exists M so that |a(u, v)| < M||u|| ||v||;
  • there exists c so that images for all u.
Show that, for each bX*, there is a unique uX such that

images

As in the previous exercise, we need to prove that there is a unique uX such that images. Let images. Existence can be proven by showing that X1 = X. Indeed, if images, then

images

Uniqueness comes from a careful use of the coercivity condition.

4.3.3 Euler–Lagrange Equations

We now study how to compute the gradient of Φ(u) on a subset X of the Hilbert space H1(Ω) using several important examples.


Example 4.3.12 (Minimization problem in one dimension)
Let σ(x) ∈ C+([a, b]) = {vC([a, b]):v > 0}. For a given fL2(a, b) and images, suppose that u* is a minimizer of the following minimization problem:

images

Then, u* satisfies

images

with the boundary conditions u*(a) = α, u*(b) = β.

 


Proof.
1. In this Hilbert space H1, the inner product and the distance (or metric) between two functions u and v are given respectively by

images

2. The goal is to investigate a minimizer u*X satisfying

images

which is equivalent to

images

where images.
3. Hence,

images

4. By the chain rule and integrating by parts,

images

This can be viewed as a directional derivative of Φ at u* in the direction v, that is, the direction where the derivative is computed.
5. Since this holds for all functions images, we have

images

The last one is the Euler–Lagrange equation of the minimization problem.

 


Example 4.3.13
Consider the curve minimization problem joining two points (a, α) and (b, β):

images

If u* is a minimizer, then

images


 


Proof.
For all images, we have

images

Since the above identity holds for all images, the minimizer u* satisfies the Euler–Lagrange equation:

images

Since

images

we have

images

Hence, the straight line joining (a, α) and (b, β) is the minimizer.

 


Example 4.3.14
Let Ω be a domain in images. Consider the minimization problem:

images

If u is a minimizer, then u satisfies the Dirichlet problem:

images


 


Proof.
For all images,

images

Since this holds for all images, we have

images


 


Example 4.3.15
Consider the minimal surface problem:

images

If a minimizer u exists, it satisfies the following nonlinear problem with the Dirichlet boundary condition:

images


 


Proof.
For all images,

images

Since this holds for all images, we have

images


 


Example 4.3.16
For a given fL2(Ω), consider the total variation minimization problem:

images

If a minimizer u exists, it satisfies the following nonlinear problem with zero Neumann boundary condition:

images


 


Proof.
For all vX,

images

Since this holds for all vX, we have

images


4.3.4 Regularity Theory and Asymptotic Analysis

We briefly discuss the regularity theory for the elliptic PDE. We know that a solution of the Laplace equation has the mean value property and therefore it is analytic (“having the ability to analyze”): if we have knowledge of all derivatives of a harmonic function at one fixed point, we can get full knowledge of the solution in its neighborhood. This mean value type property can be applied in some sense to solutions of elliptic equations. Basically, Harnack's inequality comes from a weighted mean value property. For details, please refer to the book by Gilbarg and Trudinger (2001).

Let Ω be a bounded smooth domain in images. Consider the following divergence-form elliptic operator:

4.53 4.47

where images and

images

is symmetric, bounded and positive definite. Assume that there exist positive constants c1 and c2 such that

4.54 4.48

The bounded linear form images associated with a divergence-form elliptic operator L is

4.55 4.49


Theorem 4.3.17
Let fL2(Ω), images and images. Assume that images is a weak solution of Lu = f, that is,

4.56 4.50

Then, for each Ω0 ⊂ ⊂ Ω,

4.57 4.51

where the constant C depends only on images, c0 and c1. Here by Ω0 ⊂ ⊂ Ω we mean images.

 


Proof.
Set images. Choose an open set Ω1 such that Ω0 ⊂ ⊂ Ω1 ⊂ ⊂ Ω. Take a cut-off function images such that images and images. Substituting images into (4.56) gives

4.58 4.52

From (4.54), we have

4.59 4.53

Using images, we have

4.60 4.54

Similarly, using the Schwarz inequality, we have

4.61 4.55

By taking sufficiently small images, the estimates (4.59)–(4.61) lead to

4.62 4.56

for some positive constant C1.

Now, we will estimate images. From now on, we shall use the notation x = (x1, x2, x3) = r = (x, y, z) just for simple expressions. For the estimate of images, substitute images into (4.56), where images denotes the difference quotient

images

where h is very small and images. Then,

images

We write I as

images

From the assumption of ellipticity,

images

Using the Hölder inequality and Poincaré inequality,

images

where images can be chosen arbitrarily small, images depends on images, images and images.

Then, we have the estimate

images

Since this is true for arbitrarily small h, we have the estimate

images

Here, we may use a convergence theorem in real analysis. Combining this estimate into (4.59) with an appropriate choice of images leads to (4.57).


Remark 4.3.18
For beginners in this area, we recommend starting with the equation Lu = − ∇2u = f to prove Theorem 4.3.17. With this simpler model, a much simpler computation gives the estimate (4.57). Then, it is easy to see that the use of the condition (4.54) leads to an extension to the general case.

Next, we study some important techniques for asymptotic analysis. We will use a nice simple model in the book by Chipot (2009) describing asymptotic analysis for problems in a large cylinder.

For a given f(y) ∈ C([ − a, a]), let uL and images, respectively, be solutions of

4.63 4.57

and

4.64 4.58

Then, we have the following estimates.


Theorem 4.3.19
For 0 < images < L, we have

4.65 4.59

where [Limages] is the maximal integer less than Limages.

 


Proof.
The proof of (4.65) is based on the identity

4.66 4.60

where r = (x, y) and images is a linear function such that

images

The identity (4.66) can be expressed as

4.67 4.61

Since ∇ϕimages = 0 in Ωimages,

4.68 4.62

Since |∇ϕimages| ≤ 1/(Limages),

4.69 4.63

and the left-hand side of (4.69) can be estimated by

4.70 4.64

Dividing both sides of (4.70) by images, we have

4.71 4.65

From the Poincaré inequality, the right-hand term in (4.71) can be estimated by

4.72 4.66

From (4.71) and (4.72), we have

4.73 4.67

where

images

The estimate (4.73) can be simplified as

4.74 4.68

Similar arguments for getting (4.74) with changing ϕimages give

4.75 4.69

Hence, we have

4.76 4.70

where θ = θ(1) = 4a/(1 + 4a) < 1. Now, we will estimate Φ(L). Since

images

we have

images

and, therefore,

images

This completes the proof.

We can generalize this simple asymptotic analysis (4.76) to an elliptic PDE:

images

For details, please see section 6.3 in the book by Chipot (2009).

4.4 Tikhonov Regularization and Spectral Analysis

In this section, we briefly present a functional analytic approach for the Tikhonov regularization method, which is probably the most commonly used method when the problem images is ill-posed. Regularization is an important tool to deal with an ill-posed problem by imposing a priori information on the solution. We discuss the regularization techniques for solving T*Tx = T*y, where T*T is an n × n matrix having a large condition number. For more detailed aspects, the reader may refer to the book by Engl et al. (1996).

Let X and Y be Hilbert spaces. We denote by images the set of linear operators T:XY. For each images, the Hilbert spaces X and Y, respectively, can be decomposed into

images

where

  • N(T): = {xX:Tx = 0} is the kernel of T,
  • R(T): = {Tx:xX} is the range of T,
  • N(T): = {xX:〈x, z 〉 = 0 for all zN(T)} is the orthogonal space of N(T),
  • images is the orthogonal space of R(T).

Definition 4.4.1
For each images, its generalized inverse, denoted by T, is defined as follows:
  • D(T) = R(T) ⊕ R(T),
  • T:D(T) → N(T) is a linear operator such that N(T) = R(T) and

images


 


Theorem 4.4.2
Let images and let x = Ty. Then,

images

where T* is the dual operator of T, which requiresTx, y 〉 = 〈x, T*yfor all xX and yR(T).

 


Proof.
The proof goes as follows:

images


 


Theorem 4.4.3
Let images. The generalized inverse T:D(T) → N(T) is bounded if and only if R(T) is closed, that is, images.

 


Proof.
The proof follows from the closed graph theorem in Theorem 4.4.12.

4.4.1 Overview of Tikhonov Regularization

This section is based on a lecture by Bastian von Harrach. To get insight into Tikhonov regularization, we consider the simplest operator images, which is an m × n matrix. Assume that the problem images is ill-posed. For ease of explanation, we assume that images satisfies the following:

  • images is injective and bounded,
  • images (very large),

where T* is the transpose of T. Consider the ill-posed problem:

images

or the corresponding least-squares problem

images

In practice, the data b are always contaminated by noise; hence we may consider the following problem:

images

where bδ are the noisy data. Here, the norm || · || is the standard Euclidean distance.

From the assumption that images, unregularized solutions (T*T)−1T*bdelta are prone to magnify noise in the data btrue, where btrue are the true data. Hence, (T*T)−1T*bdelta may be very different from the true solution xtrue: = (T*T)−1T*btrue even if δ ≈ 0. Hence, the goal is to find a good operator Gδ:XX so that Gδ ≈ (T*T)−1 and GδT*bδ is a reasonably good approximation of xtrue when bδbtrue.

The motivation of Tikhonov regularization is to minimize not only ||Txbδ|| but also ||x||. By adding a regularization parameter, α, we consider the following minimization problem:

4.77 4.71


Lemma 4.4.4
For α > 0, T*T + αI is invertible and its inverse is bounded by

4.78 4.72

More precisely,

4.79 4.73


 


Proof.
Using the property of the adjacent operator T*, we have

images

Hence T*T + αI is injective. We can prove that T*T + αI is surjective by the Riesz lemma or Lax–Milgram theorem. Therefore T*T + αI is invertible and

images

Next, we will prove (4.79). Using the fact that

images

(see Exercise 4.4.5), we have

images

This proves (4.79). Here, the last inequality in the above estimate comes from

images

(see Exercise 4.4.5) and the following estimate:

images


 


Exercise 4.4.5
Prove the following.
1. Prove that (T*T + αI)−1T* = T*(TT* + αI)−1 by showing the identity

images

2. Prove that T(T*T + αI)−1 = (TT* + αI)−1T.

 


Theorem 4.4.6
The minimization problem (4.77) has a unique solution, and the minimizer of (4.77) is given by

images


 


Proof.
The minimization problem (4.77) can be written as

images

which is the same as

images

The corresponding Euler–Lagrange equation is

images

Owing to the invertibility of T*T + αI in Lemma 4.4.4, the minimizer xα must satisfy xα = (T*T + αI)−1T*bδ.

To maximize the effect of the regularization, it is important to choose an appropriate regularization parameter α. The strategy is to choose an appropriate parameter α = α(δ), which can be viewed as a function of δ, such that

images

Hence, we must have α(δ) → 0 as δ → 0 because xtrue = (T*T)−1T*btrue. We need to make an appropriate choice of α(δ) so that it provides a “stable” convergence xα = (T*T + αI)−1T*bδxtrue as δ → 0.


Theorem 4.4.7
Let Txtrue = btrue and denote Gα: = (T*T + αI)−1. Let xα = GαT*bδ with ||bδbtrue|| ≤ δ. Then, for α > 0, we have

images


 


Proof.
Since xα = GαT*bδ, we have

images

where the last inequality comes from Lemma 4.4.4.

In the proof of the previous theorem for a fixed δ, note the estimate

images

The term ψ1(α): = ||GαT*(bδbtrue)|| indicates the amplified measurement error related to the difference between the exact data and the measurement data, and the term ψ2(α): = ||GαT*btruextrue||, called the regularization error, illustrates how well GαT* approximates T−1. Roughly speaking, the optimal α would be a quantity near the intersecting point of ψ1(α) = ψ2(α) as shown in Figure 4.1.

Figure 4.1 Error estimation

4.1

4.4.2 Bounded Linear Operators in Banach Space


Definition 4.4.8
A normed linear space X equipped with a norm || · || is called a Banach space if it is complete.

The following are important examples of Banach spaces in PDE theory:

  • images for images;
  • images the closure of images with norm images;
  • images.

Note that the vector space images equipped with the Lp-norm is a normed space but not a Banach space.


Theorem 4.4.9 (Baire category)
Let X be a Banach space. If images and images, then one of the Xk contains an open ball.

 


Proof.
If every Xn does not contain an open set, we can choose the following Cauchy sequences {xn} and {rn}.
1. Choose images (because images is open).
2. For n = 1, 2, …, there exists images with images (because images is open).
Since X is complete, xnxX. However, xXn for all n, which contradicts the assumption images.

 


Theorem 4.4.10 (Uniform boundedness principle or Banach–Steinhaus)
Let X be a Banach space and Y be a normed linear space. Let images. If images for all xX, then

images


 


Proof.
From the assumption, we have

images

From the Baire category theorem, there exists Xn that contains an open ball Br(x*):

images

This implies that

images


 


Theorem 4.4.11
Let T:XY be linear, where X and Y are Banach spaces. If T(X) = Y, then T is open. Moreover, if T(X) = Y and T is one-to-one, then T−1 is continuous.

 


Proof.
Denote Br: = {xX:||x|| < r} and images. From the Baire category theorem and the assumption images, there exists N such that T(BN) contains an open ball images. This proves that T is open because images for all xX and a > 0. If T is one-to-one, it is obvious that T−1 is continuous.

 


Theorem 4.4.12
Let T:XY be linear, where X and Y are Banach spaces. Then, T is bounded if and only if the graph {(x, Tx):xX} is closed with respect to the product norm ||(x, y)|| = ||x||X + ||y||Y.

 


Proof.
If the graph G: = {(x, Tx):xX} is closed, G can be viewed as a Banach space and hence the projection map Π1:GX defined by Π1((x, y)) = x has a bounded inverse due to the open mapping theorem. Since the projection map Π2:GY defined by Π2((x, y)) = y is continuous, then images is bounded. Conversely, if T is bounded, then clearly G is closed because xnx implies TxnTx.

 


Theorem 4.4.13
Let X be a Hilbert space. If an operator T:XX is linear and symmetric, then T is bounded.

 


Proof.
From the closed graph theorem, it suffices to prove that the graph G: = {(x, Tx):xX} is closed. We need to show that if xnx and Txny, then y = Tx. From the assumption that T* = T, we have

images

Since Txy is orthogonal to all ϕ ∈ X, we must have Tx = y and hence the graph G is closed.

We should note that, in order to apply the closed graph theorem, the domain of T should be a Banach space. To see this, consider the Hilbert space X = L2([0, 1]) equipped with L2-norm. We know that C2([0, 1]) is a dense subset of the Hilbert space X = L2([0, 1]). It is easy to see that the Laplace operator T:C2([0, 1]) → X defined by

images

is unbounded because

images

Defining images, we can extend the unbounded operator T to V by defining the extension operator images:VX in such a way that images whenever it makes sense. Clearly images is a closed operator. However, images is an unbounded operator, but it does not contract to the closed graph theorem because V is not complete (not a Banach space). At this point, we should note that V is a dense subset in L2([0, 1]) since images.


Theorem 4.4.14 (Hahn–Banach)
Let X be a normed linear space over images (or images) equipped with a norm || · || and let Y be a subspace of X. Let images with images(x) ≤ ||x|| for xY. Then, there exists an extension images such that images and images, where images.

 


Proof.
We will only prove the case where X is a normed linear space over images. Choose images and set images, the vector space spanned by Y and x1. From the usual induction process and Zohn's lemma, it suffices to show that there exists an extension images such that images and images(x) ≤ ||x||. Since for images we must have images for all images and yY, we only need to show the existence of images satisfying

images

or

images

Hence, the existence of images requires

images

which is equivalent to

images

Rearranging the above inequality, we need to show that

images

The above inequality holds owing to the convexity of the norm || · || and the fact that

images


 


Theorem 4.4.15 (Dual of images: Riesz representation theorem)
Let images and 1/p + 1/q = 1. For each images, we define a linear functional images by

images

Then, images. On the other hand, for each images, there exists images such that images(f) = ∫fg dx for all images.

 


Proof.
From the Hölder inequality, images. Taking images, we get images and images, which lead to images. Hence, images.
For images ∈ (Lp)*, we can define a kind of Radon–Nikodym derivative:

images

For the proof of gLq, please refer to the book by Lieb and Loss (2001).

 


Theorem 4.4.16 (Dual of Hilbert space: Riesz representation theorem)
Let H be a Hilbert space equipped with an inner product 〈 ·, · 〉 . For each images, there exists gH such that images(f) = 〈f, gfor all fH.

The proof is left as an exercise for the reader.

4.4.3 Regularization in Hilbert Space or Banach Space

Regularization is an important tool to deal with an ill-posed problem by approximating it as a well-posed problem using a priori information on solutions. We discuss regularization techniques for solving K*Kx = K*y, where K*K is an n × n matrix having a large condition number.


Definition 4.4.17
Let images.
  • The map K:XY is said to be a compact operator if images is compact, where B1 is the unit ball in X. Equivalently, K is compact if, for any bounded sequence {xn} in X, {Kxn} has a convergent subsequence.
  • The operator norm of K is images.
  • For an eigenvalue images, Xσ: = {xX:K*Kx = σx} is called a σ-eigenspace of K*K.

Throughout this section, we assume that images is a compact operator. Note that K*K maps from X to X and KK* maps from Y to Y. Both K*K and KK* are compact operators and self-adjoint, that is, 〈K*Kx, x′ 〉 = 〈x, K*Kx′ 〉 for all x, x′ ∈ X.


Exercise 4.4.18
Show that

images


 


Theorem 4.4.19
If images is a compact operator, the self-adjoint compact operator K*K:XX satisfies the following.
  • There exist eigenvalues σ1 ≥ σ2 ≥ ··· ≥ 0 and the corresponding orthonormal eigenfunctions {vn:n = 1, 2, …} of K*K such that

images

  • For each σn > 0, images and

images

up to n.
  • If K*K has infinitely many different eigenvalues, then images.
  • images.
  • Writing un = Kvn/||vn||, we have

images


 


Proof.
Since K is compact, it is easy to prove that K*K:XX is compact, that is, for any bounded sequence {xn} in X, there exists a convergent subsequence images. Hence, we have a singular system (σn;vn, un). Moreover, images; if not, K*K is not compact. The remaining proofs are elementary.

 


Theorem 4.4.20
Let K:R(K) ⊕ R(K)N(K) be a generalized inverse of a compact operator images. Then,

images

and

images

Equivalently,

images


 


Remark 4.4.21
Assume that images. Then, images and hence, for any arbitrarily small images,

images

This means that any small error in the data y may result in a large error in a reconstructed solution x = Ky.
To see this effect clearly, let us consider the backward diffusion equation. For a given time t > 0, let Kt:L2([0, π]) → L2([0, π]) be the compact operator defined by

images

where u(x, t) is a solution of the diffusion equation utuxx = 0 in images with the initial condition u(x, 0) = f(x) and the homogeneous boundary condition u(0, t) = u(π, t) = 0. The standard method of separable variables gives the explicit representation of Ktf:

images

where images and

images

Hence,

images

and its generalized inverse is

images

Though any small perturbation of f results in a small perturbation of Ktf, the generalized inverse is severely ill-posed when t is not small. To be precise, suppose that we try to find u(x, 0) from knowledge of noisy data images that is a small perturbation of the true data u(x, 2) = v1(x). Then, the computed solution

images

is very different from the true solution images. This means that the true solution u(x, 0) = e2v1(x) is negligibly small in the computed solution images.

For a compact operator K, the following are equivalent:

  • K is unbounded;
  • R(K) is not closed, that is, images;
  • images;
  • if images is a singular system for K such that images, then there are infinitely many σn > 0 such that σn → 0 as images.

Hence, if K is unbounded and K is compact, the problem Kx = y is ill-posed, that is, a small error in the data y may result in a large error in the solution x = Ky. Recall that K = (K*K)K*. Numerous regularization techniques have been used to deal with this ill-posedness.

We now introduce the Tikhonov regularization method. Let images be a compact operator. To explain the regularization method, it will be convenient to express the compact operator K*K in the form

images

where Pλ is the projection map from X to the set images. Note that images. Hence, if xD(K*K), then

images

Using the above property, we can show that

images

where images is defined by

images

For yD(K), the solution x = Ky can be expressed as

4.80 4.74

If images is compact and R(K) is not closed, then the operator K is unbounded and the problem Kx = y is ill-posed. If images, then

images

On the other hand,

images

This reconstruction method is called Tikhonov regularization. The solution

images

can also be interpreted as a minimizer of the energy functional:

4.81 4.75

Denoting images, the regularization images can be viewed as an approximation of K in the sense that

images

Here, we should note that N(K*) = R(K). The most widely used iteration method for solving y = Kx is based on the identity x = x + αK*(yKx) for some α > 0. This method is called the Landweber iteration, which expects xkx by setting

images

With this Landweber iteration, yD(K) if and only if xk converges. Indeed,

images

which leads to

images

Regularization is to replace the ill-conditioned problem Kx = y by its neighboring well-posed problem in which we impose a constraint on desired solution properties. The problem of choosing the regularization parameter α in (4.81) is critical. In MR-based medical imaging, we often need to deal with the tradeoff between spatial resolution and temporal resolution. Accelerating the MR imaging speed requires the phase encoding lines in k-space to be skipped, and results in insufficient k-space data for MR image reconstruction via inverse Fourier transformation. To deal with the missing data, one may use the minimization problem (4.81), which consists of the fidelity term (ensuring that the k-space data y are consistent with the image x) and the regularization term (which incorporates the desired solution properties). Nowadays many research areas use regularization techniques.

4.5 Basics of Real Analysis

We now briefly review the real analysis techniques that are used in this chapter. For more details, we refer to the books by Lieb and Loss (2001), Marsden (1974), Rudin (1970) and Wheeden and Zygmund (1977). Throughout this section, we assume that Ω is a bounded domain in images (n = 1, 2 or 3) and images is a function.

4.5.1 Riemann Integrability

For a quick overview of the Riemann integral, we will explain its concept only in two-dimensional space. Let images be a bounded domain and let images be a bounded function. We enclose Ω in a rectangle and extend f to the whole rectangle B = [a1, b1] × [a2, b2] by defining it to be zero outside Ω. Let images be a partition of B obtained by dividing a1 = x0 < x1 < ··· < xn = b1 and a2 = y0 < y1 < ··· < ym = b2:

images

  • Define the upper sum of f by

images

and the lower sum of f by

images

  • Define the upper integral of f on Ω by

images

and the lower integral of f on Ω by

images

We say that f is Riemann integrable or integrable if

images

If f is integrable on Ω, we denote

images

4.5.2 Measure Space

For a domain E in images, we will denote the area of the domain E by μ(E). We can measure the area μ(E) based on the area of rectangles. If Q = (a, b) × (c, d) and Qj for j = 1, 2, … is a sequence of rectangles with images for jk, then we measure them in the following ways:

images

It seems that any open domain E is measurable since any open set can be expressed as a countable union of rectangles. Then, can we measure any subset in images? For example, consider the set images, where images is the set of rational numbers. What is its area μ(E)? It is a bit complicated. Measure theory provides a systematic way to assign to each suitable subset E a positive quantity μ(E) representing its area.

The triple images is said to be a measure space if the following are true:

1. X is a set;
2. images is a σ-algebra over the set X, that is, closed under a countable union and complementations;
3. μ is a measure on images, that is,
a. μ(E) ≥ 0 for any images,
b. for all countable collections {Ej:j = 1, 2, … } of pairwise disjoint sets in images, images, and
c. μ(∅) = 0.

If the σ-algebra images includes all null sets (a null set is a set N such that μ(N) = 0), then μ is said to be complete. Let images denote the collection of all subsets of X.


Definition 4.5.1 (Outer measure)
An outer measure on a non-empty set X is a set function μ* defined on images that is non-negative, monotone and countably subadditive.

With the use of the outer measure, we can describe a general constructive procedure for obtaining the complete measure. Let images be an algebra of sets and images a set-valued function such that μpre(∅) = 0. For AX, we define

images

Then, μ* is an outer measure.


Example 4.5.2
Let images and let images be the σ-algebra generated by the set of all open rectangles in images. Define

images

Then images is a measure space but it may not be complete. This ρ is called the pre-measure. For AX, we define

images

Then, μ* is an outer measure.

The following Caratheodory definition provides a method of constructing a complete measure space images.


Definition 4.5.3
Let μ* be an outer measure on a set X. A subset AX is said to be μ*-measurable if

images


The Lebesgue measure on images is an extension of the pre-measure defined by μpre((a, b]) = ba.

4.5.3 Lebesgue-Measurable Function

Throughout this section, we restrict ourselves to the standard Lebesgue measure μ in X (images or images) that is generated by the outer measure,

images

where μpre is a pre-measure defined on open sets in X. For example, in images, μpre(U) is the volume of U.

A function images is said to be measurable if f−1(U) is measurable for every open set U.


Exercise 4.5.4
Given two functions f and g, we define

images

1. Then images is measurable if images for all images.
2. Prove that, if f and g are measurable, then so are f + g, fg, fg, fg, f+, f and |f|.
3. If {fj} is a sequence of measurable functions, then images and images are measurable.

From now on, we assume that E, Ej are measurable sets. The characteristic function of E, denoted by χE, is the function defined by

images

Let images be the set of all simple functions, a finite linear combination of characteristic functions:

4.82 4.76

Let images be the set of measurable functions and let

images

For given images, we say that f = 0 holds almost everywhere (abbreviated a.e.) if μ({x:f(x) ≠ 0}) = 0.


Theorem 4.5.5
Any images can be approximated by a sequence of simple functions

4.83 4.77

where En, k = f−1((k2n, (k + 1)2n]) and images. Moreover, each ϕn satisfies

images


The proof is straightforward by drawing a diagram of ϕn. From the above theorem, we can prove that, for any measurable function f, there is a sequence of simple functions ϕn such that ϕnf on any set on which f is bounded. The reason is that f can be decomposed into f = f+f, where f = max{f, 0} and f = max{ − f, 0}.

Now, we are ready to give the definition of the Lebesgue integral using the integral of a measurable simple function images, which is defined as

4.84 4.78

We use the convention that images. Let images be a vector space of measurable simple functions. Then the integral images (as a function of images) can be viewed as a linear functional on images, that is, the operator images is linear.


Lemma 4.5.6
Let images be a measure space. For a non-negative images and images, define

4.85 4.79

Then, images is also a measure space.

 


Proof.
It suffices to prove this for ϕ = χA where images. The proof follows from the following identity for images of mutually disjoint images:

images


Using the definition (4.84) of the integral of a measurable simple function, we can define the integral of a measurable function f ≥ 0.


Definition 4.5.7
The integral of images is defined by

4.86 4.80


From the definition, we obtain that fg implies ∫f dμ ≤ ∫g dμ.


Theorem 4.5.8 (Monotone convergence theorem)
For a non-decreasing sequence images, we have

images


 


Proof.
Since images, there exists f = limnfnMable. Since images and fnf, we have

images

It remains to prove limnfn dμ ≥ ∫f dμ. Since images, it suffices to prove that, for any α, 0 < α < 1, and any images with ϕ ≤ f,

images

Let En = {fn ≥ αϕ}. Then,

images

Since ν is a measure and images, limnν(En) = ν(X) = ∫ϕ dμ. Thus, limnfn dμ ≥ α∫ϕ dμ.

 


Exercise 4.5.9
Prove that, if images and images for some images, then

images


 


Proposition 4.5.10
Let images. Then, ∫f dμ = 0 if and only if f = 0 a.e.

 


Proof.
If images, then the statement is immediate. If f = 0 a.e. and ϕ ≤ f, then ϕ = 0 a.e., and hence images.
Conversely, if ∫f dμ = 0, then

images

and therefore f = 0 a.e. because

images


 


Lemma 4.5.11 (Fatou's lemma)
For any sequence images, we have

images


 


Proof.
Since images, we have

images


 


Definition 4.5.12
(L1(X, dμ)) A function images is integrable if images and images. We denote by L1(X, dμ) the class of all integrable functions. For fL1(X, dμ), we define

images


It is easy to see that L1(X, dμ) is a vector space equipped with the norm ||f|| = ∫|f| dμ satisfying the following:

1. ||f|| ≥ 0, for all fL1;
2. ||f|| = 0 if and only if f = 0 a.e.;
3. ||λf|| = |λ| ||f||, for all fL1 and every scalar λ;
4. ||f + g|| ≤ ||f|| + ||g||, for all f, gL1.

Now, we will prove that L1(X, dμ) is a complete normed space (or Banach space). To prove this, we need to study several convergence theorems.


Theorem 4.5.13 (Lebesgue dominated convergence theorem (LDCT))
If {fn} ⊂ L1 with fnf a.e. and there exists gL1 so that |fn| ≤ g a.e. for all n, then

images


 


Proof.
Since g + fn ≥ 0, it follows from Fatou's lemma that

images

This leads to images.
On the other hand, applying the same argument to the sequence gfn ≥ 0 yields

images


 


Exercise 4.5.14
Let

images

denote the fundamental solution of the heat equation in one dimension. Let fn(x) = G(x, 1/n). Then, fnf = 0 a.e. and

images


The above exercise explains the reason why the LDCT requires the assumption that {fn} is dominated by a fixed L1 function g.


Exercise 4.5.15
Let {fj} ⊂ L1 so that images. Show that images. Using the LDCT, prove that there exists fL1 such that

images


From (4.83), for any function fL1, there exists a sequence images such that

images

From the LDCT, ||ϕnf|| = ∫|ϕnf| dμ → 0. Hence, images is dense in L1, that is, every element in L1 is an L1-limit of a sequence of elements in images.


Exercise 4.5.16
Show that a sequence of continuous functions

images

Using this idea, show that images is dense in images.

 


Exercise 4.5.17
Let images be a mapping. Suppose that f is differentiable with respect to t and that

images

Then, F(t) = ∫f(x, t) dμ is differentiable on atb and

images

Note that, for each t ∈ (a, b), we can apply the LDCT to the sequence

images

because |hn| ≤ g from the mean value theorem.

4.5.4 Pointwise, Uniform, Norm Convergence and Convergence in Measure

Let Ω be a subdomain in images. The sequence of functions fkC(Ω) is said to converge pointwise to f if, for each images, fk(r) → f(r). We say that the sequence of functions fkC(Ω) converges uniformly to f if images.


Example 4.5.18
Consider the following.
1. The function fk(x) = xk converges pointwise to

images

on the interval [0, 1], but images for all k. Hence, the convergence is not uniform.
2. The function

images

converges uniformly to sinx in [0, 1].
3. The following functions images satisfy images in some sense:
a. images uniformly;
b. ϕn = χ(n, n+1) → 0 pointwise;
c. ϕn = nχ(0, 1/n) → 0 a.e.
d. Define a sequence images for j = 0, …, 2k − 1 and k = 0, 1, 2, …. Then, images and ||ψk, j − 0|| = 2k → 0, while images for any x.

 


Definition 4.5.19
The sequence {fn} is said to converge in measure to f if

images

and {fn} is said to be a Cauchy sequence in measure if

images


For example, the sequences ϕn = (1/n(0, n), χ(0, 1/n) and images converge to 0 in measure, but the sequence ϕn = χ(n, n+1) does not converge to 0 in measure.


Theorem 4.5.20
Suppose {fn} is a Cauchy sequence in measure. Then,
  • there exists images such that fnf in measure,
  • there exists images such that images a.e.,
  • f is uniquely determined a.e.

 


Proof.
Choose a subsequence nk such that images,

images

Then, denoting images and images, we have

images

For images and j > i, we have

images

Therefore, images exists for images. By letting f(x) = 0 for xZ, we have gkf a.e. It is easy to prove that the sequence gk converges to f uniformly on images because images. We can prove fnf in measure from

images


The statement [ fnf in L1 ] implies [ fnf in measure ] since

images

From the proof of the previous theorem, if [ fnf in L1 ], then we can choose a subsequence nk such that images,

images

Hence, according to Theorem 4.5.20, [ fnf in L1 ] implies the statement [images gkf a.e. ].


Theorem 4.5.21 (Egorov's theorem)
Let images and let fnf a.e. Then, for any images, there exists FE so that images and fnf uniformly on F.

Proof. Since fnf a.e., there exists ZE so that μ(Z) = 0 and fn(x) → f(x) for images. Hence, it suffices to prove the theorem for the case when Z = ∅.

Let images. Then, images for n = 1, 2, … . Hence, there exists mn such that images. Let images. Then, images. Moreover, if j > mn, then images Hence, fnf uniformly on F.

4.5.5 Differentiation Theory

Throughout this section, we consider a bounded function images. We will study a necessary and sufficient condition that f′ exists almost everywhere and

images

Recall that the derivative of f at x exists if all four of the following numbers have the same finite value:

images

According to Lebesgue's theorem, every monotonic function images is differentiable almost everywhere.


Example 4.5.22
If f is a Cantor function, then f′ = 0 almost everywhere but

images


To understand why images for a Cantor function f, we need to understand the concept of absolute continuity.


Definition 4.5.23
A measure ν is absolutely continuous with respect to μ if and only if images. A non-negative fBV[a, b] (the space of bounded variation functions in [a, b]) is absolutely continuous with respect to μ if and only if the measure ν(E) = ∫Efis absolutely continuous with respect to μ. The measures μ and ν are mutually singular, written μ⊥ν, if and only if

images


 


Theorem 4.5.24 (Absolute continuity)
If fexists almost everywhere, f′ ∈ L1(dμ) and

images

then f is absolutely continuous.

 


Proof.
We want to prove that, for a given images, there exists δ so that images
If |f′| is bounded, then we choose images and

images

If f′ ∈ L1(dμ) but not bounded, then we decompose

images

This is possible because the bounded functions are dense in L1(dμ). The result follows by choosing images.

From this theorem, the Cantor function is not absolutely continuous.

Every continuous and non-decreasing function f can be decomposed into the sum of an absolutely continuous function and a singular function, both monotone:

images

where g is an absolutely continuous function.


Theorem 4.5.25 (Radon–Nikodym, Riesz representation)
Let ν be a finite measure on [a, b], and suppose that ν is absolutely continuous with respect to μ. Then, there exists fL1(X, dμ) such that

images


References

Akoka S, Franconi F, Seguin F and le Pape A 2009 Radiofrequency map of an NMR coil by imaging. Magn. Reson. Imag. 11, 437–441.

Chipot M 2009 Elliptic Equations: An Introductory Course. Birkhäuser Advanced Texts. Birkhäuser, Basel.

Engl HW, Hanke M and Neubauer A 1996 Regularization of Inverse Problems. Mathematics and Its Applications. Kluwer Academic, Dordrecht.

Feldman J and Uhlmann G 2003 Inverse Problems. Lecture Note. See http://www.math.ubc.ca/∼feldman/ibook/.

Gilbarg D and Trudinger N 2001 Elliptic Partial Differential Equations of Second Order. Springer, Berlin.

Lieb EH and Loss M 2001 Analysis, 2nd edn. Graduate Studies in Mathematics, no. 14. American Mathematical Society, Providence, RI.

Marsden JE 1974 Elementary Classical Analysis. W. H. Freeman, San Francisco.

Rudin W 1970 Real and Complex Analysis. McGraw-Hill, New York.

Stollberger R and Wach P 1996 Imaging of the active B1 field in vivo. Magn. Reson. Med. 35, 246–251.

Wheeden RL and Zygmund A 1977 Measure and Integral: An Introduction to Real Analysis. Monographs and Textbooks in Pure and Applied Mathematics, vol. 43. Marcel Dekker, New York.

Further Reading

Evans LC 2010 Partial Differential Equations. Graduate Studies in Mathematics, no. 19. American Mathematical Society, Providence, RI.

Folland G 1976 Introduction to Partial Differential Equations. Princeton University Press, Princeton, NJ.

Giaquinta M 1983 Multiple Integrals in the Calculus of Variations and Non-linear Elliptic Systems. Princeton University Press, Princeton, NJ.

Grisvard P 1985 Elliptic Problems in Nonsmooth Domains. Monographs and Studies in Mathematics, no. 24. Pitman, Boston, MA.

John F 1982 Partial Differential Equations. Applied Mathematical Sciences, vol. 1. Springer, New York.

Kellogg OD 1953 Foundations of Potential Theory. Dover, New York.

Reed M and Simon B 1980 Functional Analysis. Methods of Modern Mathematical Physics, vol. I. Academic Press, San Diego, CA.

Rudin W 1970 Functional Analysis. McGraw-Hill, New York.

Zeidler E 1989 Nonlinear Functional Analysis and Its Applications. Springer, New York.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.218.70.79