Anda di halaman 1dari 19

Chapter

3 Strong and Weak Forms for one-


dimensional problems

In Chapter 3, the strong and weak forms for several physical models are
developed in one dimension. The strong form is the name given to the governing
equations and the boundary conditions. The governing equations are usually partial
differential equations, but in the one-dimensional case they become ordinary differential
equations. The weak form is an integral form of these equations which is needed to
formulate the finite element method. Differential equations in one dimension for axially
loaded elastic bars, heat conduction and electric current flow will be considered to
demonstrate the basic steps in formulating the strong and weak forms. We will also
examine the concept of various degrees of continuity, or smoothness, which will play an
important role in developing finite element methods.

3.1 The Strong Form in one-dimensional problems


3.1.2 The Strong Form for an axially loaded elastic bar

We consider the static response of an elastic bar of variable cross-section such as


shown in Figure 3.1. This is an example of a problem of linear stress analysis, where
we seek to find the stress distribution σ ( x ) in the bar. The stress will result from the
deformation of the body, which is characterized by the displacements of points in the
body, u ( x ) . The displacement results in a strain deonoted by ε ( x ) ; strain is a
dimensionless variable. As shown in Fig. 3.1, the bar is subjected to a body force or
distributed loading b ( x ) ; The body force is in units of force per volume, the distributed
axial loading is in units of force per length. For convenience and without loss of
generality the two types of loading will be expressed in units of force per unit length. In
addition loads can be prescribed at the ends where the displacement is not prescribed;
these loads are often called tractions and denoted by σ ; these loads are in units of force
per area. The body force could be due to gravity (if the bar were placed vertically instead
of as shown), a magnetic force, or a thermal stress); its units are force per unit volume.
At the two ends of the bar either a load can be specified or the displacement of the bar
can be specified. These specifications at the end points will result in the boundary
conditions of the problem.

The bar must satisfy the following conditions:

1. it must be in equilibrium
2. it must satisfy the elastic stress-strain law known as Hooke’s law σ = Eε
3. the displacement field must be compatible
4. a relationship between the strain and displacement, known as a strain-
displacement equation

The differential equation for the bar follows from the equilibrium of the forces
acting on any segment of the body. Consider an equilibrium, ∑ Fx = 0 , of an
infinitesimal element as shown in Figure 3.1.

⎛ ∆x ⎞⎟
−N (x ) + b ⎜⎜x + ⎟ ∆x + N (x + ∆x ) = 0
⎝ 2 ⎠⎟

Rearranging terms and dividing by ∆x we obtain

N (x + ∆x ) − N (x ) ⎛ ∆x ⎞⎟
+ b ⎜⎜x + ⎟= 0
∆x ⎝ 2 ⎠⎟
dN
If we take the limit of the above as ∆x → 0 , the first term is the derivative so the
dx
above can be written as

dN
+b = 0 (3.1)
dx

This is the equilibrium equation expressed in terms of the internal force N. The internal
force N is related to the stress by:

N = Aσ (3.2)

The strain-displacement (or kinematical) relation is obtained by applying the


engineering definition of strain we used in Chapter 2 for an infinitesimal segment. The
elongation is given by u ( x + ∆x ) − u ( x ) and the original length by ∆x , so

elongation u (x + ∆x ) − u (x )
ε= =
orig.length ∆x

Taking the limit of the above as ∆x → 0 , we recognize the definition of a derivative, so

2
du
ε= (3.3)
dx

Figure 3.1: Geometry and boundary conditions for the elasticity problem

Substituting Eq.(3.2) and Eq.(3.3) into Eq.(3.1) yields:

d ⎛ du ⎞
⎜AE ⎟⎟ + b = 0 ; 0 < x < l (3.4)

dx ⎝ dx ⎠⎟

In the above, u(x) is the dependent variable, which is the unknown. This differential
equation is another form of the equilibrium equation. As can be seen from how we have
constructed it, it implicitly includes the stress strain law and the definition of strain.
Compatibility is met by requiring the displacement to be continuous. More will be said
about the required continuity later.
To solve the above differential equation, we need to prescribe boundary conditions at
the two ends of the bar. For the purpose of illustration, we consider the following specific
boundary conditions: at x = 0 the force, N is prescribed, and at x = l the displacement,
u, is prescribed, i.e.:

⎛ du ⎞
N (x = 0)/ A(x = 0) = σ(x = 0) = ⎜⎜E ⎟⎟⎟ = σ
⎝ dx ⎠x =0 (3.5)
u(x = l ) = u

The governing differential equation (3.4) along with boundary conditions (3.5) is called
the Strong Form of the problem, which is summarized below:

3
d ⎛ du ⎞
⎜AE ⎟⎟ + b = 0 on 0 < x < l

dx ⎝ dx ⎠⎟
⎛ du ⎞⎟
⎜⎜E ⎟ = σ(x = 0) = σ (3.6)
⎝ dx ⎠⎟x =0
u(x = l ) = u

It should be noted that σ, u , and b are given; these is the data that describes the
problem. The unknown is the displacement u ( x ) . A more general discussion on the
boundary conditions is in given Section 3.3.

3.1.2 The Strong Form for heat conduction in one dimension


Consider a bar of length l as shown in Fig. 3.2. Heat energy is supplied to the bar,
by, for example, heating an electric wire embedded in the bar or through a surface. Heat
supply per unit length and time is denoted by Q and is measured in units of energy per
unit length and time. It is positive if the heat is added to the system. We consider a
steady state problem, i.e. a system that is not changing with time. To establish the
differential equation that governs the system, we consider heat balance (or conservation
of energy) on an infinitesmal element of the bar.

⎛ ∆x ⎞⎟
S (x ) + Q ⎜⎜x + ⎟ ∆x − S (x ++x ) = 0
⎝ 2 ⎠⎟
(3.7)
dS
or, =Q
dx

4
Figure 3.2: Geometry and boundary conditions for the heat conduction problem
where S (x ) is the total heat that enters the cross-sectional area A(x ) . Define the heat flux
q as:
S (x )
q(x ) = (3.8)
A(x )

The constitutive equation for heat flow, which relates the flux (which corresponds
to the stress in the elastic bar) to the temperature, is known as Fourier’s law (it
corresponds to Hooke’s law in elasticity):

dT
q = −k (3.9)
dx

where k is the thermal conductivity (which must be positive) and T is the temperature. A
negative sign appears in Eq. (3.9) because the heat flow direction is from high (hot) to
low temperature (cold).
Inserting Eq. (3.9) and Eq. (3.8) into Eq. (3.7) yields:

d ⎛ dT ⎞⎟
⎜Ak ⎟ +Q = 0 ; 0 < x < l (3.10)
dx ⎜⎝ dx ⎠⎟

For Ak = constant, we obtain:

d 2T
Ak +Q = 0 ; 0 < x < l (3.11)
dx 2

5
At the two ends of the bar, either the flux or the temperature must be prescribed.
This gives the boundary conditions. In this example, we consider the specific boundary
conditions q(x = 0) = q and T (x = l ) = T . The corresponding Strong Form for the
heat conduction problem is given by:

d ⎛ dT ⎞⎟
⎜Ak ⎟ +Q = 0 on 0 <x <l
dx ⎜⎝ dx ⎠⎟
dT
q = −k = q on x =0 (3.12)
dx
T = T on x = l

Further discussion on boundary conditions is given in Section 3.3.

3.1.3 Diffusion in one dimension


In the diffusion problem the unknown is the ion concentration, c. The ion flux qc
is related to the ion concentration by Fick’s law as follows:

dc
qc = −kc
dx
where kc is the diffusion coefficient. Ions move due to the gradient in ion concentration.
The balance equation for the diffusion problem is similar to that of the heat conduction
problem. The resulting differential equation is given by

d ⎛ dc ⎞
⎜Akc ⎟⎟ + Qc = 0 on 0 <x <l

dx ⎝ dx ⎠⎟

where Qc is ion source.

3.1.4 Electric flow in one dimension


The motion of an electric charge flux qV is proportional to the voltage gradient.
This is described by Ohm’s law:
dV
qV = −kV
dx

where kV is electric conductivity and V is the voltage. Again, the balance equation is
similar to that of the heat conduction problem.

d ⎛ dV ⎞⎟
⎜AkV ⎟ + QV = 0 on 0 <x <l

dx ⎝ dx ⎠⎟

where QV is electric charge source.

6
3.2 The Weak Form in one dimension
In general, the exact solution of the Strong Form of boundary value problems is
not possible for problems with complicated geometries. Therefore, most boundary value
problems are solved numerically by methods such as the the Finite Difference Method
(FDM). However, the finite difference method is not as versatile as the finite element
method in dealing with arbitrary geometries, so most engineering software has been
developed in a finite element format.
In order to apply the FEM, the partial differential equations have to be restated in
an integral form. This integral form is called the Weak Form. A weak form is an integral
form of the equations that is completely equivalent to the governing equation and certain
boundary conditions, i.e. the strong form. In many disciplines the weak form has specific
names. For example, it is called the principle of virtual work in stress analysis.
To illustrate how the Weak Form is developed, we consider the Strong Form of
the stress analysis problem given in (3.6). We start by multiplying Eq. (3.6)a by an
arbitrary function w(x ), a function called a weight function or a test function. We then
integrate the product over the problem domain which here is the interval [0, l]:
l
⎡d ⎛ du ⎞ ⎤
∫ w ⎢ ⎜⎜AE ⎟⎟⎟ + b ⎥dx = 0 ∀w
0 ⎣⎢ dx ⎝ dx ⎠ ⎦⎥

For convenience, we rewite the above in the equivalent form


l l
d ⎛ du ⎞
∫ w ⎜⎜AE ⎟⎟dx + ∫ wbdx = 0 ∀w (3.13)
0
dx ⎝ dx ⎠⎟ 0

where ∀w denotes that w(x ) is an arbitrary function, i.e. that the above has to hold for
all functions w(x ) . There are some requirements on the smoothness and boundary values
of w(x ) , but we will get to those later. It is important to note the fact that the weak form
has to hold for all functions w(x ) : without this arbitrariness of the weight function, a
weak form is not equivalent to the strong form (see Section 3.3.5).

One could use the above to develop a finite element method, but it would require
smoother approximations than are generally used (and are in fact hard to construct),
because the expression still contains second derivatives. Furthermore, the stiffness
matrix would not be symmetric. Therefore, we recast it in a form containing only first
derivatives, which will furthermore be a symmetric.
To recast the weak form so that only first derivatives appear, recall the rule for
taking the derivative of a product

d df dw df d dw
(wf ) = w +f ⇒ w = (wf ) − f
dx dx dx dx dx dx

Integrating the right hand side of the above over the domain [0 ,l], we obtain

7
l l l
df d dw
∫ w dx =
dx ∫ dx
(wf )dx − ∫ f
dx
dx
0 0 0

Now we make use of the fundamental theorem of calculus, which states that the integral
of a derivative of a function is the function itself. This enables us to replace the first
integral on the RHS by a set of boundary values
l l l
df l dw dw
∫ w dx = (wf ) 0 − ∫ f
dx dx
dx = w (l ) f (l ) − w (0) f (0) − ∫ f
dx
dx (3.14)
0 0 0

Eq. (3.14)is known as integration by parts. We will find that integration by parts useful
whenever we relate strong forms to weak forms.
du
To apply the integration by parts formulas to Eq.(3.13), let f = AE . Then
dx
Eq. (3.14) can be written as:
l l l
d ⎛ du ⎞ ⎡ du ⎤ dw du
∫ w ⎜⎜AE ⎟⎟⎟dx = ⎢wAE ⎥ − ∫ AE dx (3.15)
0
dx ⎝ dx ⎠ ⎣⎢ dx ⎦⎥ 0 0
dx dx

Using Eq. (3.15), Eq. (3.13) can be written as follows:

l l l
⎡ du ⎤ dw du
⎢wAE ⎥ −
⎢⎣ dx ⎥⎦ 0 ∫ dx
AE
dx
dx + ∫ wbdx =0 ∀w
0 0

Rearranging the terms in the above and writing out the boundary term gives
l l
dw du ⎛ du ⎞ ⎛ du ⎞
∫ AE dx = ⎜⎜wAE ⎟⎟⎟ − ⎜⎜wAE ⎟⎟⎟ + ∫ wbdx ∀w (3.16)
dx dx ⎝ dx ⎠ l ⎝ dx ⎠ 0
0 0

We will now examine the boundary terms, the first two terms on the RHS of the
above. We will see throughout the development of finite element methods that the
boundary terms must be treated carefully and well understood. If we write the boundary
du
term at x = 0 as σ = −σ ( x = 0) = − E x = 0 (i.e., positive prescribed stress σ at x = 0
dx
gives rise to compressive or negative stress), then we have
l l
dw du
∫ dx
AE
dx
dx = (wAσ ) l + (wA) 0 σ + ∫ wbdx ∀w (3.17)
0 0

The above equation is not a useful Weak Form since it involves the stress at x = l , which
is unknown. This difficulty can be circumvented by constructing a set of weight

8
functions that are arbitrary everywhere in the interval (0, l) but vanish at x = l . Then the
weak form becomes
l l
dw du
∫ dx
AE
dx
dx = (wA) 0 σ + ∫ wbdx ∀w such that w (l ) = 0 (3.18)
0 0

We will see in Section 3.3.5 that the above weak form is equivalent to the equilibrium
equation (3.6)a with stress boundary condition (3.6)b. In order for the weak form to be
completely equivalent to the strong form, we must then require the approximation
functions to satisfy the displacement boundary conditions (3.6)c. Since it is essential for
the equivalence that the approximation functions satisfy the displacement boundary
conditions, they are often called essential boundary conditions. We will see in Section
3.3.5 that the stress boundary conditions follow from the weak form and that the
approximation need not be constructed to satisfy them, so they are called natural
boundary conditions.

3.3 Generalization of the Strong and Weak Forms


3.3.1 General Strong Form for one-dimensional stress analysis
We will now consider the more general situation, where instead of specifying a stress
boundary condition at x = 0 and a displacement boundary condition at x = l ,
displacement and stress boundary conditions can be prescribed at either end. For this
purpose, we will need a more general notation for the boundaries.

The boundary of the one dimensional domain, consisting of the two points,
x = 0 and x = l , will be denoted by Γ . The portion of the boundary where the
displacements are presribed is denoted by Γu ; these as you recall, are the essential
boundary conditions. The boundary where where the stress is prescribed is denoted by
Γσ ; these are the natural boundary conditions.

The stress and displacement cannot be prescribed at the same boundary point.
Physically, this can be seen to be impossible by considering a bar such as Figure 3.1. If
we could prescribe both the displacement and the force on the right hand side, this would
mean that the deformation of the bar is independent of the applied force. It would also
mean that the material properties have no effect on the force-displacement behavior of
the bar. Obviously, this is physically unrealistic, so we require that any boundary point
be either a prescribed stress point or a prescribed displacement point. We write this
mathematically as Γσ ∩ Γu = 0 . We will see from subsequent examples that this can be
generalized to other systems: natural boundary conditions and essential boundary
conditions cannot be applied at the same boundary points.

We will often call boundaries with essential boundary conditions essential


boundaries; similarly, boundaries with natural boundary conditions will be called natural

9
boundaries. We can then say that a boundary cannot be both a natural and an essential
boundary.

It can also be shown that one type of boundary condition is needed at each
boundary point, i.e. we cannot have any boundary points at which neither an essential nor
a natural boundary condition is applied. In other words, any point of a boundary is either
an essential boundary or a natural boundary and their union is the entire boundary.
Mathematically, this can be written as Γσ ∪ Γu = Γ .

To summarize the above, at any boundary point, either an essential or a natural


boundary condition must be applied, but we can not apply both at the same boundary
point. In our shorter nomenclature, we can say that any boundary must be an essential
boundary or a natural boundary, but it can’t be both. These conditions are very important
and can be mathematically expressed by the two conditions we have stated in the above:

Γσ ∪ Γu = Γ Γ σ ∩ Γu = 0 (3.19)

Using the above terminology, we can rewrite the strong form (3.12) as

d ⎛ du ⎞
⎜AE ⎟⎟ + b = 0 0 < x < l

dx ⎝ dx ⎠⎟
du
σn = En = σ on Γσ (3.20)
dx
u = u on Γu

In the above, we have added a unit normal to the body and denoted it by n; as can
be seen from Fig. 3.1 it is given by n = −1 at x = 0 , n = +1 at x = l . This trick enables us
to write the boundary condition in terms of the forces applied at the two ends. For
example, when a positive force per unit area σ is applied at the left-hand end of the bar
in Fig. 3.1, then the stress at that end is negative, i.e. compressive, since at that end
σ n = −σ = σ . At any right hand boundary point, n = +1 and so σ n = σ = σ .

3.3.2 The Weak Form for heat conduction in one dimension


We start by multiplying Eq. (3.11) by the weight function w(x ) and integrating it
over the problem domain gives.
l l
d ⎛ dT ⎞⎟
∫ w ⎜⎜Ak
dx ⎝
⎟dx + ∫ wQdx = 0
dx ⎠⎟
∀w (3.21)
0 0

Using the integration by parts of the first term in Eq. (3.21) yields:

10
l l l
dw dT ⎡ dT ⎤
∫ dx
Ak
dx
dx = ⎢wAk
⎣⎢ dx ⎦⎥ 0 ∫0
⎥ + wQdx ∀w (3.22)
0

Recalling that the boundary term can be expressed as


l
⎡ dT ⎤ l
⎢wAk ⎥ = −[wAq ]0 = − (wAq )x =l + (wAq )x =0
⎣⎢ dx ⎦⎥ 0
= − (wAqn )x =l − (wAqn )x =0 (3.23)
= − (wAqn )Γ = − (wAq )Γq

where qn = q on Γq and w = 0 on ΓT . With this in mind and in combination with Eq.


(3.22) the Weak Form can be rearranged as follows:
l l
dw dT
∫ dx
Ak
dx
dx = − (wA) Γ q + ∫ wQdx
q
∀w such that w (ΓT ) = 0 (3.24)
0 0

Notice the similarity between Eqs. (3.24) and (3.18).

3.3.3 General Weak Form for One Dimensional Stress Analysis


We will next develop the weak form for this more general version of the strong form,
Eq.(3.20). We again multiply the strong form by the weight function and integrate over
the domain, which gives Eq. (3.15) as before
l l l
dw du ⎛ du ⎞
∫ AE dx = ⎜⎜wAE ⎟⎟⎟ + ∫ wbdx ∀w (3.25)
dx dx ⎝ dx ⎠ 0
0 0

We next consider the first term on the RHS of the above and rewrite it by using the
normal n defined above
l
⎛ du ⎞ ⎛ du ⎞ ⎛ du ⎞ ⎛ du ⎞ ⎛ du ⎞
⎜⎝⎜wAE dx ⎠⎟⎟⎟ = ⎝⎜⎜wAE dx ⎠⎟⎟⎟ − ⎝⎜⎜wAE dx ⎠⎟⎟⎟ = ⎝⎜⎜wAE dx n ⎠⎟⎟⎟ + ⎝⎜⎜wAE dx n ⎠⎟⎟⎟ (3.26)
0 l 0 l 0

In the above we have simply used the sign of the normal n to change the sign of the
second term; we will see later in two dimensional formulations that the normal appears
naturally in this form on boundary terms.

Since the far RHS of (3.94) now includes all boundary points in the same form,
we can write that

11
l
⎛ du ⎞ ⎛ du ⎞ ⎛ du ⎞ ⎛ du ⎞
⎜wAEn ⎟⎟ = ⎜⎜wAEn ⎟⎟ = ⎜⎜wAEn ⎟⎟ + ⎜⎜wAEn ⎟⎟ (3.27)
⎜⎝ ⎟
dx ⎠ 0 ⎝ ⎟
dx ⎠ Γ ⎝ ⎟
dx ⎠ Γ ⎝ dx ⎠⎟ Γ
u σ

where the second equality in the above follows from the fact that the entire boundary
consists of the displacement boundary and the stress boundary and they do not overlap,
i.e. Eq. (3.19).

In general we can state that for any one dimensional domain denoted by Ω with
boundaries Γ we have
df ( x )
∫Ω dx dx = ( fn ) Γ (3.28)

provided that f ( x ) is sufficiently smooth function. Sufficient smoothness conditions


will be discussed in Sections 3.3.3 and 3.4.4.

Substituting the (3.27) into (3.25) and denoting the integral in one dimensional domain
b

[a, b ] as ∫≡∫ gives


a Ω

dw du ⎛ du ⎞ ⎛ du ⎞
∫ AE dx = ⎜⎜wAEn ⎟⎟⎟ + ⎜⎜wAEn ⎟⎟⎟ + ∫ wAdx ∀w (3.29)
dx dx ⎝ dx ⎠ Γu ⎝ dx ⎠ Γσ
Ω Ω

Inserting (3.20)b into the first two RHS terms of (3.29), we obtain

dw du ⎛ du ⎞
∫ AE dx = ⎜⎜wAE ⎟⎟⎟ + (wAσ ) Γ + ∫ wbdx ∀w
dx dx ⎝ dx ⎠ Γu σ
Ω Ω

du
On the prescribed displacement boundary, the stress and hence is unknown.
dx
Therefore, we perform the same trick as before: we require the weight function to vanish
on the prescribed displacement boundaries. The weak form then becomes
L
dw du
∫ dx
AE
dx
dx = (wAσ ) Γ + ∫ wbdx
σ
∀w such that w = 0 on Γu
0 Ω

The above is the weak form for the general one-dimnesional stress analysis
problem (3.20). Note that it requires the weight function to vanish on the essential
boundaries. As before in Section 3.2, this weak form will not lead to satisfaction of the
essential boundary condition, so the approximation function must be required to satisfy
the essential boundary conditions. With this proviso, we can then state the weak form as:

Find a function u ( x ) such that u ( x ) = u on Γu so that

12
dw du
∫ dx
AE
dx
dx = (wAσ ) Γ + ∫ wbdx
σ
∀w such that w = 0 on Γu (3.30)
Ω Ω

3.3.4 Continuity
Although we have now developed a weak form, we still have not specified how smooth
the weight functions and approximation functions must be. Before examining this topic,
we will first define some notation for various orders of smoothness, i.e. continuity. A
function is called a C n function if its derivatives of order j, where 0 ≤ j ≤ n exist and
are continuous functions in the entire domain.

We will be concerned mainly with C 0 , C −1 and C1 functions. Examples of these


are illustrated in Figure 3.3. As can be seen, a C 0 function is piecewise continuously
differentiable, i.e. its first derivative is continuous except at selected points. The
derivative of a C 0 function is a C −1 function. So for example, if the displacement is a
C 0 function, the strain is a C −1 function. Similarly, if a temperature field is a C 0
function, the flux is a C −1 function if the conductivity is C 0 . In general, the derivative of
a C n function is C n −1 .

f(x)
C1
C −1

C0

x
a/2 a 3a/2

−1 0 1 1
Figure 3.3: Examples of C , C and C functions: C function has continuous derivatives at
x = a , but discontinuous second derivatives at x = a ; C 0 function has discontinuous
derivatives at x = a ; C −1 function is discontinuous at x = a / 2 and x = 3a / 2

3.3.5 Integrability
The smoothness that is required in the weight function and the approximation functions
depends on the integrability of the weak form: they must be smooth enough so that the
integral equation, such as Eq. (3.30), can be evaluated. Any function which is C −1 and
not singular is integrable. Since only first derivatives appear in the integral form,
C 0 continuity should then suffice for the weight and approximation functions. This
continuity requirement can also be justified physically. For example, in stress analyis, a
C −1 displacement field would be associated with gaps or overlaps at the points of

13
discontinuity of the function. This would violate the notion of compatibility of the
displacements. Although gaps are dealt with in more advanced methods to model
fracture, they are not accessible to the methods we are developing here. Similarly in heat
conduction, a C −1 temperature field would imply an infinite heat flux at selected points,
which is not physically reasonable.

We say the derivative of a function is square integrable if the internal energy, Π (u ) ,


defined as
1
( )
2
Π (u ) = ∫ du dx AEdx
2 Ω
is bounded, i.e., Π (u ) < 0 . u Π = Π (u ) is often termed as an energy norm. Note
that the internal enery is obtained by replacing the weight function by the trial function in
the first term of Eq. (3.30). The space that has square integrable derivatives is known as
the Sobolev space of degree one, denoted by H 1 . It can be proved that H 1 is a subspace
of C 0 , i.e. H 1 ⊂ C 0 . In other words, if a function has square integrable derivatives, it is
C 0 function.

An example of the function, which is C 0 , but not H 1 is depicted in Problem 3-3.


To this end we define the spaces U and U 0 for the trial and weight functions,
respectively, as
{
U = u (x ) u (x ) ∈ H 1 , u = u on Γu }
(3.31)
U0 = {w x
( ) w (x ) ∈ H ,w = 0 on Γ }
1
u

The above means that U is a set of functions that are continuous, have square integrable
derivatives and satisfy the essential boundary condition on Γu . The space U 0 is quite
similar except that it vanishes at the essential boundary.

Now we restate the weak form as follows:

find u (x ) ∈ U such that


dw du (3.32)
∫ dx
AE
dx
dx = (wA) |Γσ σ + ∫ wbdx ∀w ∈ U 0
Ω Ω

dw du
Denoting ε (w ) = and σ (u ) = AE then Eq. (3.32) can be interpreted as the
dx dx
principle of virtual work. The weight function w (x ) is called virtual displacement and
ε (w ) is the virtual strain. The left hand side represents the virtual work of internal
stresses, whereas the right hand side has the virtual work of imposed tractions and the
forces applied at the ends of the bar.

Similarly, we can state the weak form for heat conduction problems as:

14
find T (x ) ∈ U such that
dw dT (3.33)
∫ dx
Ak
dx
dx = − (wA) Γ q + ∫ wQdx
q
∀w ∈ U 0
Ω Ω

Where the spaces U and U 0 used in Eq. (3.33) are identical to those defined in Eq.
(3.31) except that u is replaced by T , u by T and ( Γu , Γσ ) by ( ΓT , Γq ).

Remarks
a) It can be seen from the above that in solving the Weak Form, we find the
particular function that satisfies the Weak Form. The set of approximation
functions from which the solution is chosen are functions that are continuous
and satisfy the essential boundary condition. The weight functions and
approximation functions must be constructed so that they satisfy these
conditions.
b) The approximation functions do not need to satisfy the natural boundary
conditions. As we shall see, the natural boundary conditions are naturally met
by solving the Weak Form. It can be seen that the highest order derivative
appearing in the Weak Form is one as opposed to two in the Strong Form, and
therefore the function need not be as smooth as when Eq. (3.13) is used as a
weak form. Eq. (3.13) is still integrable-how do we finesse that?
c) Functions w(x ) and u(x ) appear symmetrically in the first integral in
Eq.(3.32), whereas they do not in Eq. (3.13). In Eq. (3.32) both the
approximation and weight functions appear as first derivatives, whereas in the
first integral in Eq. (3.13), the test function appears directly whereas the trial
function appears as a second derivative. It will be seen that consequently, Eq.
(3.32) leads to a symmetric stiffness matrix and a set of symmetric linear
algebraic equations.

3.3.6 The Equivalence between Weak and Strong Forms


In the previous sections we constructed the Weak Form from the Strong Form. To show
the equivalence between the two, we will now prove that the Weak Form implies the
Strong Form. The equivalence of the two will insure us that when we solve the weak
form, then we have a solution to the strong form.

du
Let v = AE and apply integration by parts to the left side of Eq.(3.32):
dx
l l
dw du ⎛ du ⎞ d ⎛ du ⎞
∫ AE dx = ⎜⎜wnAE ⎟⎟⎟ − ∫ w ⎜⎜AE ⎟⎟⎟dx
dx dx ⎝ dx ⎠ Γ dx ⎝ dx ⎠
0 0

15
l
⎛ du ⎞ d ⎛ du ⎞
= ⎜⎜wnAE ⎟⎟⎟ − ∫ w ⎜⎜AE ⎟⎟⎟dx
⎝ dx ⎠ Γσ dx ⎝ dx ⎠
0

where in the second line Γ has been replaced by Γσ since w = 0 on Γu and Γσ is the
complement of Γu ; see Eq. (3.19). Substituting the above into Eq. (3.33) and placing
the integral terms on the left hand side and the boundary terms on the right hand side
gives
l
d ⎛ du ⎞ ⎛ du ⎞
∫ w dx ⎜⎜⎝AE dx + b⎠⎟⎟⎟dx = wA ⎝⎜⎜En dx − σ ⎠⎟⎟⎟ Γσ
(3.34)
0

The key point in the proof is that w(x ) is an arbitrary function. It can be assumed to be
anything we need in order to prove the equivalence.

First, we define
⎡d ⎛ du ⎞ ⎤
w = ψ(x ) ⎢ ⎜⎜AE ⎟⎟⎟ + b ⎥ (3.35)
⎢⎣dx ⎝ dx ⎠ ⎥⎦

where ψ is smooth and positive ψ(x ) > 0 on 0 < x < l ; and vanish on the boundary
ψ(0) = ψ(1) = 0 . An example of a function satifying the above requirements is
ψ(x ) = x (l − x ) .
Eq. (3.35) implies that w(x = 0) = w(x = l ) = 0 . Inserting(3.35) into Eq. (3.34)
yields:
l 2
⎡d ⎛ du ⎞ ⎤
∫ ψ ⎢ ⎜⎜AE ⎟⎟⎟ + b ⎥ dx = 0 (3.36)
⎢⎣dx ⎝ dx ⎠ ⎥⎦
0 

integrand

Note that the integrand in Eq. (3.36) is positive at every point in the problem domain. So
the only way the integral will vanish is if the integrand is zero at every point. Hence

d ⎛ du ⎞
⎜AE ⎟⎟ + b = 0 0 <x <l (3.37)

dx ⎝ dx ⎠⎟

This is precisely the differential equation appearing in the Strong Form Eq. (3.6)a. Now
after substituting Eq. (3.37) into Eq. (3.34) we are left with:

⎛ du ⎞
wA ⎜⎜En − σ ⎟⎟⎟ = 0 ∀w ∈ U 0 (3.38)
⎝ dx ⎠Γ
σ

Since the weight function is arbitrary on Γσ , we select it such that:

16
w Γσ ≠ 0

Substituting above into Eq. (3.38) and recalling that the cross-sectional area A(x) > 0
implies:
du
nE = σ on Γσ
dx

and thus we recover the natural boundary condition Eq.(3.6)b.

Problems Chapter 3
Problem 3-1: Given the Strong Form representing the heat conduction problem in a
circular plate:
d ⎛ dT ⎞⎟
k ⎜r ⎟ + rQ = 0 0<r ≤R
dr ⎜⎝ dr ⎠⎟

dT
Natural B.C.: (r = 0) = 0
dr

Essential B.C: T (r = R ) = 0

Where R is the total radius of the plate, Q is the heat supply per unit length along the
plate radius, T is the temperature and k is the conductivity. Assume that k, Q and R are
given:

a) Solve the differential equation with the boundary conditions and show that the
temperature distribution along the radius is given as:

Q
T = (R2 − r 2 )
4k

b) Construct the weak form for the heat conduction problem in a circular plate with the
essential and natural boundary conditions given above.

Problem 3-2: Given the Strong Form for the circular bar in torsion:

17
m(x )

φ (x = 0) = φ x

M (x = l ) = h

x =0 x =l

Figure 3.4: Definition of the torsion problem

d ⎛ d φ ⎞⎟
⎜JG ⎟ + m = 0 0 ≤ x ≤ l
dx ⎜⎝ dx ⎠⎟

⎛ dφ ⎞
Natural B.C.: M (x = l ) = ⎜⎜JG ⎟⎟⎟ = M
⎝ dx ⎠x =l

Essential B.C: φ (x = 0) = φ

where m(x) is a distributed moment per unit length, M is a torsion moment, φ is an angle
of rotation. G is a shear modulus, J is a polar moment of inertia given by:
πC 4
J = and C is the radius of the circular shaft.
2
a) Construct the weak form for the circular bar in torsion.

b) Assume that m(x) = 0 and integrate the differential equation given above. Find the

integration constants using boundary conditions.

Problem 3-3: Consider a problem on 0 ≤ x ≤ l which has a solution in the form of


⎪ λ
⎛ 1 ⎞⎟ x l

⎪ − ⎜⎜ ⎟ x≤

⎪ ⎝2⎠ l 2
u =⎨

⎪⎛ x 1 ⎞⎟
λ λ
⎛ 1 ⎞⎟ x l
⎪⎜ ⎜
⎜ − 2 ⎠⎟ − ⎝⎜ 2 ⎠⎟ l x>
⎪⎜
⎩⎝ l
⎪ 2

0
a) Show that for λ > 0 the solution u is C in the interval 0 ≤ x ≤ l
1
b) Show that for 0 < λ ≤ 1/ 2 the solution u is not in H

18
Problem 3-4: Consider an elastic bar with a variable distributed spring p(x ) along its
length as shown in Figure 3.5.

p(x)
x
σ

Figure 3.5: Elastic bar with distributed springs

Assume the bar length l , cross-sectional area A(x ) , Young modulus E (x ) , distributed
loading b(x ) and boundary conditions as shown in Figure 3.5.

a) Construct the strong form


b) Construct the weak form

Problem 3-5: Consider an elastic bar in Figure 3.1. The bar is heated with temperature
T (x ) . In case of thermal loading the stress-strain law is given by
σ(x ) = E (x ) (ε(x ) − α(x )T (x ))

where α is the coefficient of thermal expansion, which may be a function of x .

a) Construct the strong form


b) Construct the weak form

Problem 3-6: Find the weak form for the following strong form:
d 2u
κ 2
− λu + 2 x 2 = 0 , κ, λ are constants, 0 < x <1
dx
Subject to u(0) = 1, u(1) = −2

19

Anda mungkin juga menyukai