Purdue University

EAS 657

Geophysical Inverse Theory

Robert L. Nowack

Lecture 14

 

Continuous Differential Operators

 

 

            We start by defining the inner product

 

 

where the bar above denotes the complex conjugate of the function.  We now use the definition of the adjoint to provide a relation between A and A* (where A* is the adjoint operator).  Thus,

 

 

First, we will discuss compatibility relations.  Let

 

                                                then                                           (1a)

and,

                                               then                                         (1b)

 

Now from the relation for the adjoint, then

 

                                                                                                                         (2)

 

since .  This is called the compatibility relation between the primary problem and the adjoint problem.  Thus, for all u and y satisfying equations (1a) and (1b), equation (2) must be satisfied.

 

            Ex)  Let  is in the null space of A*, or from the adjoint theorems, y is in Rperp(A).  Then , which is what y being in Rperp(A) signifies.

 

            Before we go ahead and look at some Green’s functions, we need to define several adjoint operators.  For differential operators, our inner products require the specification of boundary values on the volume of interest.  Thus,

 

 

The volume v and its boundary S for a physical problem is shown below where the outward normal of the boundary is .

 

 

 

 

The equation above can be used to derive A* and the relevant boundary terms.  The adjoint operator A* is defined above to make an exact differential of the term .

 

Ex)  Consider the del operator  in one dimension for real functions, then .  Now,  where b.t. signifies the boundary term.  From the definition of the adjoint,

 

 

Recall the formula for interpretation by parts,

 

In the adjoint formula above, let

 

 

and

 

 

then,

 

 

We see that

 

 

Thus, the adjoint A* makes the integrand of an exact differential.  Note that  is not self adjoint since A*  A and also the boundary term only equals zero if  or , or a mixed combination.

 

            Ex)  In 3-D, this extends to , , where  by the del operator.  Then,

 

 

and the adjoint of the del operator acting on a scalar field is

 

 

Note, we will use the term Hermitian adjoint if all i’s are changed to –i’s in the adjoint.

 

            Ex)  A = C a constant.  For this case, A* =  and the b.t. = 0.

 

Thus, A  A* for a complex number, but A = A* for a real number.

 

            Ex)    in 1-D.  For this case,

 

 

Use integration by parts in the form,

 

 

In the adjoint formula, let

 

 

and

 

 

then,

 

 

or

 

 

Thus,

 

 

and the adjoint can be written,

 

 

For this case,

 

            Ex)  1-D wave operator (the Helmhotz equation)   with   where v is the wave speed and  is the frequency.  Then,

 

 

where this is formally self adjoint.

 

            Ex)  3-D Laplacian

 

Since the following equation is true

 

 

then,

 

 

As a summary, here are some adjoints for some simple differential operators (assuming a real inner product)

 

Au

A*y

Boundary Term

 

where  is the outward unit normal on the boundary.  For complex inner products, then we need to change y to  in the boundary term.

 

            Homogeneous boundary conditions occur when the boundary term is made to vanish by choosing the boundary conditions of the original problem and adjoint problem such that the boundary term is zero for the definition of the adjoint.

 

            For the moment, we won’t comment on inhomogeneous boundary conditions except to note that we can always separate our problem  into , where

 

 

and

 

 

 

            Let’s return to the case with a special solution g when the source term , then.  We can also define solutions for the adjoint problem with

 

,

 

where x is the receiver location and x1 is the source location.  g(x,x1) is called the Green’s function for A and G(x,x1) is the Green’s function for A*.

 

            Assume a complex inner product

 

 

and the definition of the adjoint as

 

 

where b.t. signifies the boundary terms.  Then,

 

 

or,

 

                                                                                     (3a)

 

where  is the complex conjugate of the adjoint Green’s function for A*.  For homogeneous boundary conditions, the boundary term is zero.  This can be interpreted as finding u(x1) by putting a delta function source at each receiver and evaluating at all interior points using G instead of g, where g is the ordinary Green’s function for A.

 

            Ex)  For the  operator, then

 

 ,

 

assuming  on the boundary.  Note that assuming both u and  on S raises questions of incompatibility, which we will ignore for now.

 

            Now let,

 

 

where g is the ordinary Green’s function and x1 is the receiver location and x2 is the source location..  Now  and from Eqn. (3a) above for homogeneous boundary conditions where the boundary term is zero, then

 

 

This is the general reciprocity relationship between the Green’s function for A and the adjoint Green’s function for A* where the bar signifies a complex conjugate.  Thus, the ordinary Green’s function is related to the one for A* with the source and receiver interchanged and all i’s changed to -i’s.

 

            Now for self-adjoint problems,

 

 

then,

 

 

which is the reciprocity relation for the self-adjoint case.

 

For the general case, the representation theorem above can also be written assuming homogeneous boundary conditions

 

                                                                                                  (3b)

 

This form (3b) is often not as useful as (3a) since the sources  in the medium could be at many points, but there might be only one receiver position x1.

 

            Ex)  As an alternative test to determine self adjointness, one can look at the relation .  As an example,

 

 

The Green’s function for  is

 

 .

 

For this case,

 

 .

 

Thus,  is self adjoint.

 

            Ex)   is formally self adjoint for the 1-D Helmholtz operator.  The Green’s function can be written as,

 

 

Since,

 

 

But, the Helmholtz equation was shown earlier to be formally self adjoint.  This would imply that .  Thus, from the analysis of the Green’s functions, the Helmholtz equation is not strictly self adjoint even though .

 

In the figure below,  describes a wave exploding from x2 to x1.

 

 

 

 

In the next figure,  describes a wave imploding from from x2 to x1.

 

 

 

 

The sign of i shows that in the first case the source is exploding from x2 to x1 and in the adjoint case the source is imploding from x2 to  x1.

 

            Ex)  For the acoustic wave equation

 

 

Fourier transforming this from time to frequency and then from position to wavenumber gives

 

 

and

 

 

For a medium with constant velocity c (this would be a homogeneous medium not to be confused with homogeneous boundary conditions), then the Green’s function g can be written in different domains as

 

 

(x,t) Domain

(x,) Domain

(k,) Domain



1-D

- sign is right going

+ sign is left going

+ sign is right going

- sign is left going

 





2-D

where

- sign is out going
+ sign is in going

For , then

 





3-D

- sign is out going

+ sign is in going

+ sign is out going
- sign is in going

 

 

 

where  is the Dirac delta function,  is the Heaviside step function and the functions  and  are Hankel functions.  For the wave equation in the frequency and space domains, then , which is strictly not self adjoint, but still has spatial reciprocity.

 

 

 

Vector Green’s Functions

 

            Assume we have a function in three directions at each point x

 

 

For example, in a linear, elastic Earth, the wave equation in the frequency domain can be written as an operator A acting on a vector wavefield with the right hand side being the source field.  The adjoint problem can be written in terms of the adjoint operator A*.

 

 

where u is the wavefield for A and  is the source field for A, and  is the wavefield for A* and  is the source field for A* or

 

  and 

 

We define the inner product  to be

 

 

and the adjoint is obtained from

 

 

assuming homogeneous boundary conditions.

 

From  and , then

 

                                                                 (4)

 

Let  be a point force in the “i” direction at position x1, then

 

 

where j is the receiver component, i is the source direction, x is the receiver location and x1 is the source location.   is the Kronecker delta equal to 1 if i=j and zero otherwise.   is the Dirac delta function which is only nonzero if x = x1.  Then,

                                with homogeneous B.C.                    (5)

 

Now let,

 

 

then,

 

 

From equation (5), this results in the vector reciprocity relation,

 

 

            General reciprocity then involves 1) interchange source and receiver locations, 2) interchange source orientation and receiver orientation, and 3) complex conjugate.  For a self adjoint operator A* = A, then g = G and,

 

 

Note that the elastic wave equation in the frequency domain is not strictly self adjoint since for the homogeneous Green’s function

 

 

 

 

Applications to Inverse Problems

 

            Assume we are given a forward problem

 

                                                                                                                       (6)

 

where this assumes the scalar wavefield u0(x) and the source function .  We would like to to solve the following problems.

 

I)    The inverse source problem.  What would the new wavefield be if we perturbed the source function?  Then,

 

 

where, for example, the source might represent an earthquake source.  Since , then,

 

 

We can formally invert this equation using the Green’s function formalism as

 

                                                                                           (7a)

 

or

 

                                                                                               (7b)

 

by general reciprocity.  Note that for the scalar wave equation in the frequency domain (the Helmholtz equation)  and for this case

 

                                                                                            (7c)

 

Note that this is in the form of a linear inverse problem where the observed residual field  is used to invert for the source perturbation .  Eqn. (7c) is often a more useful form since  implies putting a source at each receiver point x1 and evaluating the field at all interior points x.  Assuming the number of receivers is small, this may be more computationally efficient in terms of the number of forward problems required.

 

 

II)  The inverse medium problem.  What would the new wavefield  be if the operator A is perturbed?  This might result from changing the velocity model incorporated in the operator A for the scalar wave equation.

 

            In the frequency domain, the scalar wave equation can be written in terms of the Helmholtz operator as

 

 

where v(x) is the medium velocity at all points x in the interior of the volume.  Now,

 

 .

 

We want to solve,

 

 

where  is the known source and we know u0(x) for the unperturbed problem

 

 

The perturbed problem can be expanded as

 

 

Since , this reduces to

 

 

If we assume that  are “small”, then as a first approximation we can neglect the second order term .  Then, the linearized problem for the perturbed wavefield can be written

 

 

This is called the Born approximation.  From the Green’s function formalism

 

                                                                                  (7b¢)

 

and

 

                                                                                  (7c¢)

 

where  is the perturbed field at the receiver xr and  is the original wavefield.

 

            For the Helmholtz operator, then

 

 

 .

 

Assuming

 

 

then,

 

 

Since for the scalar wave equation  and assuming a point source where

 

 

then

 

 

This reduces to,

 

                                           (8)

 

where  can be interpreted as the observed wavefield minus the calculated wavefield.   is the back propagation from the receiver locations to the interior points in the model,  is the propagation from the sources to the interior model points and  is the perturbed velocity to be solved for.  The figure below illustrates the forward and back propagated fields for a buried target zone in the Earth.

 

 

 

 

We would then have to do this for all sources and receivers to include all the data.

 

            We know how to solve this type of problem in the discrete case.  Let , then

 

 

where  is the generalized inverse.  But, for the continuous case, we need to investigate this further.  It is a linearized system for  given  where

 

 

and

 

 

We now would like to see how this is related to what reflection seismologists call seismic imaging and migration.

 

            From equation (8), the sensitivity  will only be “large” when  and  are in “phase”.  This is sometimes referred to as an imaging principle.  The reflector exists when the downgoing forward waves from the source coincide with the back propagated waves from “pseudo” sources located at the receiver locations.

 

            If  is a constant, then  will be the Green’s function for the homogeneous reference model.  The Green’s functions in equation (8) contain no “cross talk” in the perturbations.  Hence the Born approximation in equation (8) is a single scattering approximation.  It does not include multiples scattering in the perturbations.  The figure below illustrates this.

 

 

 

 

 

Seismic Imaging and Migration

 

            There are a number of steps involved in seismic imaging.  These include

 

1)  Data collection.  Seismic data is always collected in “common shot gathers”.

 

 

 

 

For multiple sources, the data can be regrouped into “common midpoint gathers”.

 

 

 

 

For a plane layered velocity structure, this would give one scattering point called the common depth point.  This is a very logical way to organize the data.  However, for laterally varying media, the scattering points will not be at the midpoint between shots and receivers.  For a simple layer over a half space, this is illustrated below.

 

 

 

 

In this case, the scattering points are all at the midpoint, where h is the half offset between the shot and receiver and xM is the midpoint location.

 

 

2)   Normal Moveout and Stack

 

            First, we need to organize the data into common midpoint (CMP) gathers at each midpoint location xM.

 

 

 

 

In order to obtain a representative reflection trace for that midpoint position, one can perform a “normal” moveout to align the arrivals and stack.  This is called a CMP stack.

 

 

 

 

If the average velocity v1 to the reflector is chosen correctly for the normal moveout correction, this stacking operation will give a large increase in coherent signal to noise.  A  increase in signal to noise is possible, where N is the number of traces stacked (all “statics” due to a near the surface weathering zone, topography, and lateral velocity variations must be removed in order to get the full benefit of the stack).

 

 

3)   Plot all the CMP stacked traces at the midpoint positions xM.  This results in a zero offset stacked seismic section in (x,t) and is illustrated below.

 

 

 

 

Of course we would need to stack the data to obtain the effective zero offset section.  Alternatively, we could just plot a near offset trace (or any fixed offset trace at xM) if we choose not to stack, but this would not have the noise cancellation benefits of the stack  This entire procedure can be viewed as a partial solution to the inverse problem since a blurred version of the “reflected” parts of the data is displayed.

 

            The zero offset section can be viewed as an ensemble of experiments performed using a moving zero-offset source-receiver pair at each position along the surface x.

 

 

 

 

Naturally, this type of reflection or echo experiment will be most sensitive to abrupt changes in the velocity.

 

            Since for normal incidence the reflection coefficient is proportional to

 

 ,

 

we will be inverting for the seismic reflectivity R and from this estimating changes in the impedance I which is the combined variable times v.  At the outset, we will assume that the smooth variations in velocity are known (say  or a constant v0).  We need this in any case to take advantage of the stack, but we could again have used just a near offset trace to get a non-stacked common offset section.  To obtain the smooth part of the velocity model, velocity analysis can be performed by maximizing the semblance or power in the stack for different trial velocity models .

 

            We now want to investigate abrupt changes in the velocity giving rise to reflections.  Thus, we separate the velocity model into a slowly varying background velocity and a fast varying reflectivity. 

 

            Now, what would the reflectivity response be to a small point scatter in the earth?  Assume initially a homogeneous background velocity model.  Also, assume a zero offset source receiver pair moving along the surface.  As shown in the figure below, a set of diffracted arrivals can be obtained from the multiple source-receiver pairs.

 

 

 

 

The 3-D point source response in the time domain is,

 

 

The received signal will then be

 

 

where D is a diffraction coefficient related to the perturbation, the point reflectivity, and * signifies convolution.  Then,

 

 

If we are just interested in timing (ignoring amplitudes), then we could consider our multiple experiment replaced by one experiment with a source placed at the scattering point in the earth and all velocities changed to (v/2).

 

 

 

 

This has an incorrect amplitude, but the correct timing, and is called the “exploding reflector model”.  The simultaneous exploding of multiple scatterers in the subsurface is illustrated below.

 

 

 

 

Thus, we replace many numerical experiments by one experiment where all reflectors are made up of point scatterers that “explode” at t = 0. 

 

            Now let’s look alternatively in the (x,t) domain.  The figure below shows one pulse originating from a semicircle of scatterers with pulses all arriving at the same time.

 

 

 

 

A semicircle of point reflectors would give one pulse in the (x-t) zero offset section at time  and location .

 

We thus have two ways of converting a zero offset section into an approximate model of the reflectivity in (x,z).

 

1)   Assuming an average velocity v, sum hyperbolas on the (x-t) record sections and put the energy in (x,z) at the point  (Figure A1).

 

2)   Take each point in the (x,t) record section and smear this energy out into semicircles in (x,z) of equal amplitude and put it at center  with radius  (Figure A2).

 

Either of these procedures is known as Kirchhoff migration in the time domain.  A commercial diffraction stack algorithm would typically work on the transformed (x,) section in the frequency domain and do the equivalent operation.  Note that if  constant, then ray tracing must be performed to find the times to the point diffractors in the subsurface.  A major problem occurs in complicated geologic structures where we don’t know either the reflectivity D(x,z) and the smooth velocity v(x,z).  There are errors in both determinations and possible tradeoffs.

 

A reference for Kirchhoff migration is:  Schneider (1978), Integral formulation for migration in two and three dimensions, Geophysics, v. 43, p. 49-76.

 

            The figure below shows an example of a CMP stacked section above and a migrated seismic section below.  The seismic line is 11 km long across the Japan Trench.  The stack section shows many hyperbola from point scatterers in the model that could be misinterpreted.  In the migration image, the diffraction pattern for each point scatterer is migrated to its apex.

 

 

 

 

 

Born Inversion

 

            To investigate seismic migration as an inverse problem, we will start with the scalar wave equation

 

 

where u(t,x) is pressure field in t and x, (t,x) is the source, and  is the spatially varying velocity.  In the frequency domain, the wave operator is

 

 

Assuming homogeneous B.C. and the ordinary Green’s function, then

 

 

The initial forward problem for operator A0 can be written

 

 

and we want to find the new field such that

 

 

Assume that the source is known and , then

 

 

or

 

 

Then, with

 

 

This is an exact equation for .  If  is a small and second order, then we can neglect it resulting in

 

 

This is called the Born approximation.  For our case,

 

 

From the Green’s function formalism

 

 

and from the special reciprocity for the wave equation, then

 

 

            In a homogeneous medium, the Green’s function can be written in the frequency domain as

 

 

For a point source at xs and a source time function spectra , then

 

 

The perturbed pressure field can then be written as,

 

                                                (9)

 

Eqn. (9) is the linearized Born approximation for the perturbed field for a slight perturbation in the velocity .  Thus  + scattered waves from  + scattered waves from  + … .  This is also called the single scattering approximation.  In doing this, we have left out all the “cross talk” between  and .  This “crosstalk” is called multiple scattering in the perturbations and is included in the  term which we left out. 

 

Considering only the linearized approximation

 

 

where B is the linearized Born operator.

 

            The inverse Born problem is to find  from .  Thus, symbolically we can write the solution as

 

 

We would be finished except that most inverse operators are hard to find. But, adjoint operators are usually easy to find.  Inverse operators can also be constructed from adjoint operators.  The adjoint operator B* takes us from the data space to the model space.

 

For the overdetermined case,         

For the underdetermined case,       

 

Since B is only an approximate, linearized sensitivity operator, we may not need to know  exactly.  Thus, let

 

 

where  is the observed wavefield at the receivers minus the calculated wavefield in the initial velocity model and  is a small number.  Now define B* from

 

 

where  is the inner problem over sources, receivers, and frequency.

 

            Define the inner products as

 

 

and

 

 

Now from the definition of adjoints above

 

, 

 

since velocity is defined with a real inner product.  Then,

 

 

and

 

 

 

 

Now,

 

 

then,

 

 

 

 

Then,

 

 .

 

            For coincident sources and receivers,

 

 

            Now, this can be transformed into the time domain where we use the following identities where F.T. denotes the Fourier Transform.

 

 

 

 

Let,

 

 

Now,

 

 

where “*” signifies convolution.  This gives,

 

 

This reduces to

 

                                                               (10a)

 

with

 

                                                                                 (10b)

 

Eqn. (10b) is the cross correlation with the second derivative of the source time function.  Eqn. (10a) says to sum energy along hyperbolas weighted by .  This is not quite the same as the empirically derived Kirchhoff formula, but we haven’t accounted for (B*B)-1 as yet.

 

            Thus, the action of adjoint B* is equivalent to a matched filter operation with the source function plus a Kirchhoff style summation.

 

            A complete velocity inversion could be posed in terms of an iterative sequence of Kirchhoff migrations.  This approach was described by Tarantola (1987).  An approximate (B*B)-1 compensates for the amplitude part to give a modified amplitude term.

 

            More complete discussions of Born inversion of seismic reflection data are described below.

 

Bleistein, N., K. Cohen, and F. Hagin (1987).  Two and one half dimensional inversion with an arbitrary reference, Geophysics, 52, 26-36.

 

Bleistein, N., K. Cohen, and J. Stockwell (2001).  Mathematics of Multichannel Seismic Imaging Migration and Inversion, Springer, New York.

 

Jin, S., R. Madariaga, J. Virieux, and G. Lambare (1992).  Two-dimensional asymptotic iterative elastic inversion, Geophys. J. Int., 108, 575-588.

 

Tarantola, A. (1987).  Inverse Problem Theory, Elsevier.