You are on page 1of 19

Chapter 2

Cylindrical Bending of a Thin Slab

2.1 Derivation of the equation

In this part we analyze the deformation of a thin slab under a distributed load q(x) and lateral compression
P. In order to find a simple analytical formulation we assume that the slab is thin, it has small deformations
and it is of infinite length in the out of plane direction (cylindrical deformation and loading). It is possible
to find analytical solutions also for finite homogeneous thin slab and you can find it in many engineering
books. Under the assumption of small deformation it is also possible to find a close analytical formulation
for thick slab but for our scope to understand how the elastic lithosphere can support a load and does
deforms the cylindrical bending formulation is sufficient.

Let’s take a thin slab of thickness h and length L such that h  L and very long (infinite) in the out of
plane direction (z, Figure 2.1).

Figure 2.1:

Note that the very long dimension in the z direction is often indicate as 2D plane strain condition and
imply that all the load and moments are not point quantities but are defined as quantity per unit length.
The slab is deformed in the y direction of the amount w(x). Also the deformation of the slab is small
compared to the length (w  L).

Assuming the slab to be in equilibrium (it is a problem of elastostatic), we can derive the equation relating
the slab deformation to the applied forces and material properties using a force balance.

8
Figure 2.2:

Following the notation of Figure 2.2 let’s make the force balance for a small section of length dx of the
slab. The slab is loaded by a diffuse vertical load q(x) and by lateral horizontal pressure P . The horizontal
force balance will give:

P + P + dP = 0 =⇒ dP = 0 =⇒ P = constant (2.1)

On the vertical direction we will have:

q(x) dx + V + dV − V = 0 =⇒ q(x) dx + dV = 0 (2.2)

This can be rewritten as:


dV
= −q(x) (2.3)
dx

Note that the force is upward on the left side and downward on the right side. This is due to the fact that
on the right side the little piece of the slab reacts to the remaining plate at its right that is holding it up
while on the left side is my little piece that is holding the up the remaining part of the plate. Note also
that while q(x) is the external load applied to the plate, V is an internal force within the plate!

Since the plate is bending a balance of the forces is not enough to define the status of equilibrium. Every
little piece of the slab is also feeling some torque around an axis in the z direction. On the left side the
remaining plate will impose to our little piece of plate a moment equal to M while the torque of our little
piece on the remaining part of the slab to the right will be M + dM . Again as in the case of V the 2
moments at the sides of the little piece are oriented in the opposite direction. This imply that the total
moment given by the remaining plate is:

−M + M + dM = dM (2.4)

9
Figure 2.3:

Given the deflection of the plate equal to w on the left side and w + dw on the right side, the forcing P
will create a moment (see Figure 2.1)
−P dw (2.5)

Similarly the internal force V will give a moment

V dx (2.6)

Where V is assumed to be the same at the 2 sides of the little piece of plate being dv a perturbation of
the second order (dx) is small! and thus negligible.

Summing [2.4], [2.5], and [2.6] we have:

dM − P dw = V dx (2.7)

or
dM dw
=V +P (2.8)
dx dx

If we differentiate a second time and we substitute [2.3] we get:

d2 M d2 w
= −q(x) + P (2.9)
dx2 dx2

In this equation we still have the internal torque and if we want to have a relation between the external
loads P and q(x) and the slab deformation w(x) we need to be able to express the internal moment M as
function of the deformation. Here is where the linear elasticity and the thin slab will enter in the equation.

10
Figure 2.4:

If the plate is deflected downward (Figure 2.1) the upper half is in compression and the lower half is in
extension. This means that there must be a plane in the middle that is neither stretched nor compressed.
We call this plane the Neutral Plane. On the lateral surface of the little piece (with a normal directed in
the x direction) I will have a stress acting in the same direction that will be compressive in the upper part
and extensional in the lower part. In term of normal stress σxx (called also Fiber Stress) it is positive
in the upper part, zero on the neutral plane (y=0) and negative in the lower half. This variation of the
normal stress σxx will create the bending moment such that:

ˆ +h
2
M= σxx ydy (2.10)
−h
2

Using the linear elasticity constitutive laws we can relate the stress to the deformation (strain). For isotropic
homogeneous material we have
σij = λεkk δij + 2µεij (2.11)

where λ and µ are the Lame’s parameters that describe the elastic properties of the plate. Introducing
the Young’s modulus and Poisson ratio and inverting the equations we find the expression of the strain in
function of the stresses:

1 ν ν
εxx = E σxx − E σyy − E σzz
1 ν ν
εyy = E σyy − E σxx − E σzz
1 ν ν
εzz = E σzz − E σxx − E σyy
1
(2.12)
εxy = 2µ σxy
1
εxz = 2µ σxz
1
εyz = 2µ σyz

For the plane strain assumption εzz = 0 at the same time the fact that the slab is very thin and is free to
deform in the y direction (free surface) we can include a plane stress assumption such that σxy = σyz =
σyy = 0. This reduce the set of equations [2.12] to:


 ε = E1 σxx − Eν σzz
 xx


εyy = − Eν σzz − Eν σxx (2.13)


0 = − ν σ + 1 σ

E xx E zz

11
That gives σzz = νσxx or
E
σxx = εxx (2.14)
1 − ν2

Introducing [2.14] in [2.10] we get


ˆ h
E 2
M= εxx y dy (2.15)
1 − ν2 −h
2

εxx is dependent by the distance of the neutral plane and by the radius of curvature R as in Figure 2.1.

Figure 2.5:

The length of the segment of plate measured along the mid plane by the definition of neutral plane remain
L
of the given length L and by definition of radians we have φ = R or

1 φ
= (2.16)
R L
∆L
for small angle φ, L = dx. On the other hand we know that εxx = L and from Figure 2.1we can see
that
L = Rφ
∆L = yφ (2.17)
∆L yφ y
εxx = L = Rφ = R

Implicit assumption w is small and the plate is still almost planar!!)

12
Figure 2.6:

With some geometrical construction as in Figure 2.1 we see that for the bent plate the angle φ is equivalent
to minus the angle of the tangent to the plate with the horizontal α. But since w(x) gives the geometry of
dw
the bent plate, the angle α is nothing else than the derivative of w with respect to x α = dx that means:

d2 w
 
dα d dw
φ = −dα = − dx = − dx = − (2.18)
dx dx dx dx2

1 φ φ
but from [2.16] R = L = dx that means
1 d2 w
=− 2 (2.19)
R dx

the second derivative of the flexure is related to the radius of curvature (not a big surprise!!).

Now combining [2.17] and [2.19] we get an expression of the strain in function of the deflection w:

y d2 w
εxx = =− 2y (2.20)
R dx

When I substitute [2.20] in [2.15] we get the expression of the Moment in term of the deflection:

ˆ h ˆ h
2 E d2 w 2 E d2 w 2 E h3 d2 w
M= − 2 2
y dy = − y 2 dy = − (2.21)
−h
2
1 − ν dx 1 − ν 2 dx2 −h
2
12 (1 − ν ) dx2
2

We can introduce a parameter only dependent on the plate property called flexural rigidity

E h3
D= (2.22)
12 (1 − ν 2 )

and we can rewrite the equation [2.21] as

d2 w D
M = −D 2
=− (2.23)
dx R

13
The flexural rigidity indicate how much the plate would resist to the bending when we apply a torque. The
radius of curvature of the plate would be proportional to the applied moment and its flexural rigidity.

When we substitute [2.23] in to [2.9] we get an expression relating the external loads, the flexural rigidity
and the deformation of the plate:

d4 w d2 w
D = q(x) − P (2.24)
dx4 dx2

2.2 Some engineering applications

2.2.1 Embedded plate under applied moment

Now that we have obtained the equation relating applied moments and load to the plate deformations and
internal moment and shear forces, let’s try to apply it to some example relevant to the engineering case.
Remember that we are applying the equation to cylindrical bending that means or plate is a thins slab that
is very long in the out of plane direction and that the out of plane deformation is zero.

Figure 2.7:

As first example we take a slab that is embedded at one end and subject to an applied torque at the
other as shown in Fig 2.2.1. For simplicity we assume that the plate is weigthless that imply q(x) = 0. To
compute the deformation of the slab we need to solve equation [2.24], a 4th order differential equation that
need 4 boundary conditions. In reality, it is simpler to use [2.3], [2.8], and [2.23] to reduce the equation of
a system of 2 first order equations plus a second order equation (note we still need 4 boundary conditions!)

From [2.3] we know that this imply that the shear stress on a section of the plate must be constant and
since we do not apply any force to the plate (first bc!) we can conclude that V (x) = 0.

On the other hand from [2.8] and the fact that P = 0 we know that

dM
=0 (2.25)
dx

that means that M is constant: every point of the slab independently by x feels the same torque. By
boundary conditions we know that at the end of the slab (x=L) the torque is equal to the apply one
(second bc). This means that M = Ma .

14
Figure 2.8:

Figure 2.8 gives the force balance for the problem we are studying.

From equation [2.23] we have:


d2 w
Ma = −D (2.26)
dx2

that integrated twice with respect to x give:

1
w=− Ma x2 + c1 x + c2 (2.27)
2D

The embedded boundary conditions means that at x=0 both w and its first derivative must be equal to
0. These 2 boundary conditions allow us to compute c1 = c2 = 0 and to get the searched solution of the
deformation of the embedded slab when we apply a torque to the free end:

M 2
w=− x (2.28)
2D

Clearly for a given torque the deformation of the slab is larger longer is the plate and smaller is the flexural
rigidity that is what we should expect from the force and moment balance. It is also worthy to point out
that since M is constant, in order to hold the plate the walls where the plate is embedded need to apply a
force opposite to Ma as indicated in figure 2.8.

2.2.2 Embedded slab under applied vertical load

As second example we compute the deformation of a slab under a concentrated load applied at the free
end (figure 2.10). As before our slab is very long in the out of plane direction and the load must be seen
as a linear load. We can think at this problem as a balcony that is loaded by a line of flower pots at its
end.

Figure 2.9:

Following the previous example we reduce our 4th order equation [2.24] to a system of 2 first order equations
plus a second order one. In order to achieve the force and moment balance (figure 2.10) we can use the

15
equations [2.3], [2.8], and [2.23] with the boundary conditions that V (x = L) = Va , M (x = L) = 0 (since
it is a free end it can deform in such a way that the local moment is equal to 0), plus the 2 “embedded”
dw
boundary conditions w(x = 0) = dx (x = 0) = 0.

Figure 2.10:

From equation [2.3] we get that V=constant. Since at x=L V = Va we conclude that the shear force on
the slab is constant and equal to the applied shear at the free end: V (x) = Va .
dM
From equation [2.8] we have dx = V + P dw
dx = Va . Integrated with respect to x and imposing P=0 we
get
M = Va x + c2 (2.29)

Imposing M (x = L) = 0 give us the equation:

M = Va (x − L) (2.30)

Note that M (0) = −Va L, thus the moment necessary to hold the slab embedded is Va L, that is not a big
surprise, it is equal to the force applied to the end time the length of the slab!

We can now use equation [2.23] with the value of M we have just computed and write:

d2 w d2 w Va
−D 2
= V a (x − L) =⇒ 2
= − (x − L) (2.31)
dx dx D

Integrated twice is giving:

Va 3 L 2
w=− x + x + c2 x + c3 (2.32)
6D 2D

From the “embedded” bc we get that the 2 constant are equal to 0 and our final result is:

V a x2  x
w= L− (2.33)
2D 3

2.3 Application to the lithosphere

Up to now the slab was free to move up and down without to interact with the surrounding material.
In particular we had the implicit assumption that the movement of the slab was in equilibrium with the
surrounding. In the case that we are bending a slab on air the assumption is pretty good but if we have

16
material of different density above and below our slab we need to take in consideration the fact that we
are displacing material bringing our system out of the isostatic equilibrium. For the lithosphere this is the
case. Displacing high density mantle material and substituting it with lighter thus buoyant material from
the crust we have an isostatic force that push upward.

Figure 2.11:

Let’s take the case of oceanic lithosphere deflected downward (figure 2.11 left). Let’s assume to have
a bathymetry hw , an elastic oceanic lithosphere of thickness h, and a fluid inviscid mantle half-space
underneath. If the density of the water is ρw and the density of the mantle + oceanic lithosphere is ρm ,
in the undeformed case the pressure at the depth w within the fluid half-space would be Pundef ormed =
g [ρw hw + ρm (h + w)]. If a vertical load deflect the seafloor downward of the amount w, the basin
will be filled by water and the new pressure at the top of the fluid half-space will now be Pdef ormed =
g [ρw (hw + w) + ρm h]. Since in the fluid material the pressure at the given depth must be constant
(otherwise the it would generate a flow that tend to re equilibrate it), below the deflected elastic layer I need
to have an upward pressure equivalent to the difference between the undeformed and deformed pressure.
This pressure is equivalent to Piso = g (ρm − ρw ) w. In the case of continental lithosphere (figure 2.11
right) we can assume that the basin will be filled with sediments of continental crust density. The isostatic
pressure would be in this case Piso = g (ρm = ρc ) w. If the lithosphere is loaded by the external load q(x)
(positive downward for the derivation we did 2 weeks ago), the addition of the isostatic pressure will give
an equivalent vertical load equal to q 0 (x) = qa (x) − g (ρm − ρw ) w or q 0 (x) = qa (x) − g (ρm − ρc ) w for
the two cases. When this expression is substitute in [2.24] we get the formula for the deflection of the
lithosphere including the isostatic forces that can be expressed as:

4 2
D ddxw4 + P ddxw2 + g (ρm − ρw ) w = qa (x)
or (2.34)
4
d2 w
D ddxw4 +P dx2 + g (ρm − ρc ) w = qa (x)

Note that the isostatic term can be seen as a dumper!

Let’s try to see how does the lithosphere would deform under some particular loads.

2.3.1 Stability of Earth lithosphere under end load

Often when one analyze the deformation of the elastic lithosphere assume the load P to be equal to 0.
What would happen if we apply a horizontal force P? We can see that if P exceed a critical value, an

17
infinitely long plate (L −→ ∞) will become unstable and deflect in the sinusoidal shape shown in figure
2.12.

Figure 2.12:

Let’s start from equation [2.34] in the case of oceanic lithosphere (historically the first time this study was
done to figure out if the deformation of the Indian Ocean could have been related to buckling) and assume
that q(x) = 0.

The equation can be rewritten as:

d4 w d2 w
D 4
+ P 2 + g (ρm − ρw ) w = 0 (2.35)
dx dx



This equation can be satisfied by a sinusoidal deflection w = a sin λ x if

 4  2
2π 2π
D −P + g (ρm − ρw ) = 0 (2.36)
λ λ

(derived by substitution in equation [2.35]).

Or in other words if λ is such that


2 p
P 2 − 4Dg (ρm − ρw )

2π P±
= (2.37)
λ 2D

Since λ must be real it is necessary that

P 2 − 4Dg (ρm − ρw ) > 0 (2.38)

The plate is deflected only if P is larger than a critical value


s
p Eh3 g (ρm − ρw )
Pc = 4Dg (ρm − ρw ) = = σc h (2.39)
3 (1 − ν 2 )

where σc is the critical stress associated with the force P c (remember that the force P is per unit length!).
Solving [2.39] for the critical stress we obtain
s
Eh (ρm − ρw ) g
σc = (2.40)
3 (1 − ν 2 )

18
If P is smaller than P c the plate is stable and will not buckle. As soon as the horizontal force reaches the
critical value the plate will start to buckle with a wavelength λ given by
s
4
Eh3
λc = 2π (2.41)
12 (1 − ν 2 ) (ρ m − ρw ) g

It is interesting that from an engineering point of view the critical stress is the stress that will lead a
structure to failure.
kg kg
For a typical lithosphere of E = 70Gpa; ν = 0.25; ρm = 3300 m3 ; ρw = 1000 m3 ; h = 25km, σ c =

4GP a = 40kbar. It is interesting to note that normal rocks would break under load of few hundred to
thousand bar while the normal stress drop in an earthquake is of the order of 10 to 100 bar. This means that
before to reach the buckling stress the lithosphere would fail by faulting. Thus, the creation of synclines
and anticlines by horizontal stress of the full lithosphere can not be explained as buckling of an elastic plate
and other processes (as plastic behavior) must take place on long time scale. The fact that horizontal
stresses necessary to deform the slab are so high, means that these horizontal forces would contribute only
to internal stress and not to deformation. This is why they are often set to 0 in the computation of the
elastic bending. It is also interesting to note that while buckling can not take place on the full lithospheric
scale, it is possible for thin layer of sediments. If for example we compute the critical stress in case of
E=20GPa and h=1km the critical stress is only of the order of the kbar.

2.3.2 Lithospheric deflection under periodic load

We started all this analysis of the elastic bending of plates from the discussion of isostasy. The problem was
that to apply Airy or Pratt isostatic corrections we need to have a very weak lithosphere (that is contrary to
the definition of lithosphere) and we should have much larger vertical offset that are not observed. Starting
from the analysis of gravity anomaly, Meinesz suggested that the isostasy is on a regional and not local
scale assuming that the strength of the elastic lithosphere can support part of the load.

To study how the elastic lithosphere can support the load we try to analyze the deformation we would
obtain under a periodic load (let’s think at it as a mountain range followed by a basin and so on). If we
assume a topography expressed as T = T0 sin 2π

λ x the load on the elastic slab can be thought as
 

qa (x) = gρc T0 sin x (2.42)
λ

that substituted in equation [2.24] with P=0 give the 4th order differential equation

d4 w
 

D + (ρm − ρc ) gw = ρc gT0 sin x (2.43)
dx4 λ

We can assume that the deflection of the thin elastic slab is also a periodic function with the same period
 

w = w0 sin x (2.44)
λ

19
Substituting equation [2.44] in [2.43] we can see that it is a solution of our differential equation if
 4

D w0 + g (ρm − ρc ) w0 = ρc gT0 (2.45)
λ

or
T0
w0 = (2.46)
ρm D 2π 4

ρc −1+ ρc g λ

q
D
Note that 4
ρc g has the dimension of a length, it is dependent only by the characteristic of the slab (not
the load) and it is proportional to what we call the natural wavelength of the flexure of the plate. We can
compare this natural flexure with the q
wavelength of the topographic load. If λ is very small compared to
the natural flexure wavelength, λ  ρD4
cg
the denominator of [2.46] will be very big thus the deflection
w0 will be very small (in the limit λ −→ 0 =⇒ w0 −→ 0). Very short wavelength topography can be
completely supported by lithosphere elastic strength!

q
D
If λ  4
ρc g the last term of the denominator would be pretty small and the denominator would be
completely controlled by the first 2 terms in the limit of λ −→ ∞ equation [2.46] reduce to

−→ ρc
w0 w∞ = T0 (2.47)
λ→∞ 0 ρm − ρc

That is exactly the formula to compute the root of a given topography T0 in the Airy isostatic compensa-
tion!!!

Thus, long wavelength loads are not supported by the lithosphere elastic strength but by isostatic compen-
sation!

Let’s define a variable C called degree of compensation

w0 ρm − ρc
C= = (2.48)
w0∞ ρm − ρc + D 2π 4

g λ

C will be equal to 1 in the case of a topography fully compensated by Airy’s isostasy and equal to 0 in
the case that the elastic strength of the lithosphere can support the load and it will be completely out
of isostatic equilibrium. In the first case the Free Air gravity anomaly would be almost zero while in the
second case it would be the Bouger gravity anomaly to be very close to 0 (figure).

20
Figure 2.13:

q
1 D
If I define the natural wavelength of the flexure of the plate the value 2π
4
g(ρm −ρc) it is possible to see
that loads with such λ would be half compensated (C=0.5).

Figure 2.14 give the degree of compensation as function of the load wavelength expressed as multiple of
the natural wavelength of the flexure.

Figure 2.14:

2.3.3 Bending of the elastic lithosphere under a localized linear load

Let’s now compute the deformation of the lithosphere under a linear localized load. This can be similar
to what we should observe when we look at the deformation surrounding the Hawaiian seamountain chain
(figure 2.15).

21
Figure 2.15:

In reality the Hawaiian chain is not really a localized load having a wavelength of approximately 150 km
(so as we will see compatible with the deformation) but as first approximation it is working pretty well.
Let’s assume to have an infinite slab and that we apply a load V a at x=0. As we have dome previously
we can assume that qa (x) = 0 x 6= 0 and P=0. The equation [2.34] is thus written as

d4 w
D + g (ρm − ρw ) w = 0 (2.49)
dx4

A general solution of this equation is

w = c1 ep1 x + c2 ep2 x + c3 ep3 x + c4 ep4 x (2.50)

where pi are the 4 solutions of the associate equation

Dp4 + g (ρm − ρw ) = 0 (2.51)

note that g (ρm − ρw ) is positive. This means that the equation we need to solve is
r r
4 g (ρm − ρw ) √ 4 g (ρm − ρw ) √ 4 g (ρm − ρw )
p =− =⇒ p = ± 4 −1 =± i (2.52)
D D D

√ √
2
if we remember that i= 2 (1 ± i) we get that the solutions are
r
4 g (ρm − ρw )
p = ± (1 ± i) (2.53)
4D

if we introduce the parameter α called flexural parameter

22
s
4
4D
α= (2.54)
g (ρm − ρw )

we have that pi = ± (1 ± i) αthus the general solution is

1+i 1+i 1−i 1−i


w = c1 e α + c2 e − α + c3 e α + c4 e− α (2.55)

eix +e−ix eix −e−ix


If we remember that 2 = cos(x) and 2i = sin(x) the equation [2.55] can be rewritten

x
 x  x  x
 x  x 
w = e α c01 cos + c02 sin + e− α c03 cos + c04 sin (2.56)
α α α α

Given the symmetry of the problem with respect to x=0, we can analyze only the positive part. Similar
considerations can be done for the negative part but the symmetry of the problem tell us that our solution
must be symmetric with respect to the origin.
The first requirement for the solution is that it is finite. This tell us that when x → ∞ w → 0 that
gives c01 = c02 = 0. Since we analyze only the positive part of the equation we need to apply a symmetry
boundary condition. This imply that dw
dx = 0 at x=0. This boundary conditions imply that c03 = c04 . These
2 boundary conditions imply that the equation [2.56] reduce to

x
 x  x 
w = c e α cos + sin x≥0 (2.57)
α α

3
From the equation [2.8] and [2.23] we get that D ddxw3 = V that at x=0 means

d3 w

1 4c
Va = D 3 =D 3 (2.58)
2 dx x=0 α

(Note the 1/2 in front to Va due to the symmetry!) that gives

Va α3
c= (2.59)
8D

Thus the solution for the deflection is

Va α 3 x  x  x 
w= e α cos + sin x≥0 (2.60)
8D α α

Note that the solution for a linear load is periodic with wavelength and dumping dependent on the flexural
parameter. This means that a vertical load would create a basin (depression) bounded by two flexural high
(periferal bulges or forebulge). (this make sense if you think at the motion of the fluid underneath the
elastic slab). In reality the dumping is such that in general only the first high is significant.
The amount of deflection of the depocenter of the basin is dependent on the flexural parameter and the
applied load:

Va α3
wmax = (2.61)
8D

23
From equation [??] we can compute the halfwidth of the depression x0 that is given by


x0 = α arctan (−1) = α (2.62)
4

The distance of the forebulge is given by setting the derivative of [??] equal to 0 thus:

dw 2wmax − x x
=− e α sin =0 (2.63)
dx α α

that gives
xb = α arcsin (0) = πα (2.64)

Note that both [2.62] and [2.64] are not dependent on the load. Increasing the load would increase the
amplitude of both the depth of the depocenter and the high of the forebulge but would not increase the
width of the basin.

Since the distance of the forebulge from the depocenter is easy to measure (both from gravity and
bathymetry) we can use this value to compute the regional apparent elastic thickness

3 3g (ρm − ρw ) 1 − ν 2  xb 4
h = (2.65)
E π

In the exercises of tomorrow you will compute the elastic thickness of the lithosphere supporting the
Hawaiian chain.

2.3.4 Bending at subduction

In a similar way than what we have done in the previous example, we are able to compute the deformation
of a semi infinite slab that is bent from a load or a moment applied at the free end. This is the typical
representation of a subduction zone (Figure 2.16)

Figure 2.16:

In the case that the applied moment M0 is equal to 0, the problem is very similar to the previous one. The
major difference is that we do not need to apply a symmetry boundary conditions at x=0 (thus the applied
load is the full V0 and not one half and the first derivative of the deflection does not need anymore to be
equal to 0) but we need to impose that that end of the slab is a free end thus M=0 (form equation [2.23]).

24
In a similar way of what we have done before we would get that

x
x
w = w0 e− α cos (2.66)
α

V0 α3
with α expressed by [2.54], and w0 = 4D . It is clear that also in this case we would have a periferal bulge
π 3π
and an intersection of the deflection with the zero given by: x0 = 2α and xb = 4 α. From topography
and gravity measurement it is possible to measure the distance xb − x0 thus compute the flexural parameter
also without to know the forced applied at the free end. In the case of both torque and load applied at the
free end the solution is more complex but still present a form very similar to [2.66] and the presence of x0
and xb only dependent by the flexural parameter. It is clear that the presence of the periferal bulge give a
dynamic topography supported by the flexural rigidity than can be identified as a positive gravity anomaly
offshore of a trench in a subduction region. It is also interesting to note that the the upward bending of
the slab in the periferal bulge induce some extensional stresses. This is compatible with the fact that often
offshore of the trench some normal fault earthquake are observed as shown in figure 2.17 in the Auletian
islands.

Figure 2.17:

25
Figure 2.18:

Figure 2.18 give 2 bathymetric profiles across 2 trenches and the profile calculated using the previous
assumptions. It is interesting to see that while it is working very well for the first one, for the Tonga trench
it fail to explain the deformation. The most probable explanation is that the Tonga trench bend is too
sharp and violate the small deformation assumption we made at the beginning. The deformation is so
strong that the plate can not behave anymore as an elastic material but will have some plastic behavior.
In this region, an elasto-plastic model can do a better job in interpreting the observed deformation.

26

You might also like