Sei sulla pagina 1di 41

Colorado School of Mines Department of Civil and Environmental Engineering Advanced Structural Analysis Set of Notes #1

1. Basics of Shear and Bending Moment Diagrams.


The basics of Shear force (V) and Bending Moment (M) diagrams have been taught in the elementary structural analysis course. These concepts are briefly repeated here: 1. To create V and M diagrams, a sign convention is needed. This sign convention is illustrated in Figure 1. A way to interpret the shear force convention is as follows: The shear force V is positive if the shear force is upwards on a right section surface. The bending moment M is positive if it puts the bottom side in tension. 2. To calculate the shear force and bending moment at any cross-section, we have to consider an imaginary section of the beam at that location and examine the equilibrium of one of the two cut beam pieces. Let us follow the example of Figure 2. To evaluate the shear force and bending moment at a cross-section at a distance x > 10 m from the left end, we consider a section at x that cuts the beam into two pieces and we examine the equilibrium of either piece (left or right). Either equilibrium will result in exactly the same values of shear force and bending moment. Thus, the choice of which cut part will be examined for equilibrium depends strictly on convenience. Note that in either case, the unknown values of shear V(x) and moment M(x) are

Figure 1: Sign convention for Shear Force and Bending Moment Diagrams.

Figure 2: Example Problem for M and V calculations.

In the example shown here, the equilibrium of the left part is as follows: ( ) ( ) ( )

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 1

Similarly, we can examine the equilibrium of the right part of the beam, which as will be seen is simpler since it involves fewer forces: ( ) ( ) ( )

It becomes clear that the result is the same no matter which part of the beam is examined for equilibrium. To calculate the bending moment at the same cross-section we have to examine the equilibrium of the moments about any point. Usually, the cross-section itself is the most convenient point because it removes the contribution of Vx. Equilibrium of the left section of the beam gives us:
( ) ( ) ( ) ( ) ( )

Similarly, equilibrium of the right section of the beam gives:


( ) ( ) ( ) ( )

Again, we can see, that equilibrium of either section results in the same value for the moment. 3. Let us revisit the shear equilibrium equation for the left section of the beam: ( ) ( ) ( ) It is clear from this equation that the shear at the point x is the sum of all the shearing forces to the left of the cross-section with the upward forces having a positive contribution and the downward forces having a negative contribution. Similarly, if we examine the equilibrium of the right section: ( ) ( ) ( ) We can see that the shear at the point x is the sum of all the shearing forces to the right of the cross-section with the downward forces having a positive contribution and the upward forces having a negative contribution. Similarly, the equilibrium equations for the moment indicate that the moment at a cross-section is equal to the sum of all the moments about the cross-section, to the left of the cross-section, with the clockwise moments giving a positive contribution. Also, the moment at a cross-section is equal to the sum of all the moments to the right of the cross-section about the cross-section with the counter clockwise moments giving a positive contribution. The italicized observations in item 3 will be used to produce shear force and bending moment diagrams for any beam.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 2

1.1 Example
Produce and draw the shear force and bending moment diagram of the beam shown here. Solution 1. Find reactions By satisfying the equilibrium equations, it is easy to find that the reactions are RA = 800 kN and RB = 400 kN. The horizontal reaction at A is zero.
Figure 3: Example Beam

2. Create Shear Diagram

a) (x 6m) Consider an arbitrary point at a distance x from the left end. As long as x # 6 m the actual value of x does not change the free body equilibrium. Thus, the shear at this point is the sum of all the shearing forces to the left of x which result in: ( ) Similarly ( ) or ( ) (Quadratic function) (Linear function)

Note that within this range (x 6m), the shear changes sign from positive to negative. At x = 4m V=0, which, of course is the place of maximum moment value. b) (x > 6m) Consider now an arbitrary point at a distance x > 6 m from the left end. The point is located at a distance of 9-x from the right end. Thus, the shear at this point is the sum of all the shearing forces to the right of x (more convenient to consider this, rather than the left of x): ( ) (Constant Function)

Similarly, for the moment: ( ) ( ) OR ( ) (Linear function)

The shear force and bending moment diagrams are plotted in Figure 4.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 3

Figure 4: Free body diagram, Shear Force Diagram, Bending Moment Diagram

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 4

2. More Basics of Shear and Bending Moment Diagrams: M-V Relation.


The basic relations between distributed load w, shear force V and bending moment M are: and in differential form OR and in integral form.

If we examine these equations we can conclude the following: 1. For an unloaded part of the beam (w=0) the shear force diagram is constant, while the bending moment diagram is linear. 2. For a part of the beam with uniform load (w = constant) the shear force diagram is linear, while the bending moment diagram is quadratic. 3. For a part of the beam with trapezoidal load (w = linear) the shear force diagram is quadratic, while the bending moment diagram is cubic. Considering local equilibrium we can add the following conclusions: 4. A change in the magnitude of a uniformly distributed load results in a kink in the shear force diagram. No change in the moment diagram magnitude or slope exists at that point. 5. For a concentrated force there is a jump in the shear force diagram (this is the only circumstance to have a jump in a shear diagram) equal to the concentrated force. The effect in the bending moment diagram is a kink (again this is the only circumstance to have a kink in the moment diagram). 6. For a concentrated moment there is a jump in the bending moment diagram (this is the only circumstance to have a jump in the M diagram) while there is no effect whatsoever in the shear diagram. 7. The shear force and bending moment diagram MUST satisfy the boundary conditions. Thus, a free body diagram should be drawn and then the shear force and bending moment diagram values at the ends should be equal to the applied shear forces and bending moments externally applied to those ends. Figure 5: Example of boundary condition compliance for M and V diagrams

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 5

3. More Basics of Shear and Bending Moment Diagrams: Superposition.


The principle of superposition is a fundamental principle this is applied to linear mechanics. The term linear mechanics implies that the material behavior is linear (Hooks Law) and that the deformations are small. The principle of superposition simply states that the solution of a linear problem subjected to multiple simultaneous excitations is equal to the sum of the solutions when each of these excitations is applied alone. This principle can be used significantly to our advantage by taking complex problems and analyze them as the sum of simpler components. The three MOST COMMON types of structural loads are the uniformly distributed load, the concentrated force, and the end moments applied on simply supported beams. These problems are presented in Figure 6 along with their solutions in terms of reactions, shear diagrams, and bending moment diagrams. IT IS EXPECTED THAT THE STUDENTS OF THIS CLASS WILL MEMORIZE THESE SOLUTIONS, SO THAT THEY CAN USE THEM DIRECTLY RATHER THAN REPRODUCE THEM IN THE COUNTLESS TIMES THAT THEY SHALL ENCOUNTER THEM IN THIS CLASS, AS WELL AS IN THEIR PROFESSIONAL CAREER. One of the most significant combinations of loads is presented in Figure 7. The details of this solution will be discussed in the classroom. THIS SOLUTION MUST ALSO BE MEMORIZED. Two examples, where the moment diagrams (and implicitly the shear diagrams) can be produced with little effort are presented in Figure 8. Additional examples will be discussed in the classroom.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 6

Figure 6: Fundamental Problems

Figure 7: Solution of a fundamental problem.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 7

Figure 8: Example problems where solutions of bending moment diagrams (and implicitly shear diagrams) can be achieved with little effort

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 8

4. Basic Energy Issues


4.1 Real Energy
Consider an infinitesimal element of a beam or column with length dx (Figure 9). The element is subjected to internal axial forces N, shear forces V, and bending moments M. The element is stretched by , distorted by and flexed by an angle due to these internal forces. It is easier to understand these deformations and their energy implications if we examine them using the principle of superposition, i.e. we examine their effects when they are applied one force at a time.

Figure 9: Infinitesimal element (in the x direction) subjected to axial, shear, and bending loads and deformations

Due to the application of the axial force N, the element is stretched by a length . Given the relations: , and , where A is the beam cross-sectional area, we calculate the corresponding energy as:

Due to the application of the shear force V, the element is distorted by a shear strain , or a deformation . Given that and , we calculate the corresponding energy as:

Due to the application of the bending moment M, the element is flexed by an angle . Note that and thus, = y . Thus, , where is the deflection curvature. For elastic beams, thus, we can calculate the corresponding energy as:

, and

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 9

The total energy of the infinitesimal element can now be evaluated:

The total energy of the entire beam or column can be evaluated by integration of Equation (1) over the entire length of the structural element:

(1)

U is called the internal energy of the beam or column. In beam and frame problems, the N and V energy contributions are typically very small compared to the M energy contribution and may be ignored. In truss problems, only N contributes in the total energy since V, and M are either very small or zero.

It is a basic principle of energy that the internal and external energy of a system is the same. Consider the example shown in Figure 10. It is loaded by a concentrated force P at the mid-point of the beam. We would like to calculate the deflection at the point of load application. We note that the external energy , that is, the energy due to the external forces (in this case ) is equal to Figure 10: Example of deflection calculation using the total energy of the system. .

The reactions are also external forces (easy to see when you use a free body diagram), but do not produce energy since the supports do not move. The internal energy is equal to . Note that N=0 and the term is ignored.

The moment equation is expressed by a piece-wise function as follows:

{ ( Thus, the internal energy of this beam is: )

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 10

))

Due to the symmetry of the problem it is easier to define the internal energy as two times the energy from 0 to L/2. That is: ( )

Since

, we have

or

This may appear to be an impressive approach to calculate deflections. HOWEVER, it cannot work of the external energy is anything other than .

For example, how can we find the deflection at a point other than under the applied load? That deformation does not appear in our external energy. Similarly, how do we find the deformation at a point, where there is more than one load? For example, if we have two concentrated forces And , then the external work is . How can we find ? The problem, of course is

worse, if the external load is distributed.

4.2 Virtual Work


The virtual work method solves this problem very easily. Consider the spring problem demonstrated in Figure 11. First we apply a force which results in a deformation . The work of this process is An additional force deformation due to is is then applied to the spring (total force . as these symbols are ). The resulting , while the total deformation is ( ) ( )

The total work is now equal to defined in Figure 11.

We are especially interested in the partial work Physically is the work performed on due to the deformation caused by (i.e. ). Note here that

work does not have the factor because the force is already fully developed as it is carried by the deformation . This is a very interesting, although not trivial to understand concept, which is very useful in calculating deflections of structures.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 11

Figure 11: Infinitesimal element (in the x direction) subjected to axial, shear, and bending loads and deformations

We must remember that any external load produces work which must be equaled by the corresponding internal work resulting from internal forces and deformations as was explained in 4.1. Let us consider the simple structure of Figure 12 which carries a uniformly distributed load over its girder BC. We would like to calculate the horizontal deformation of point C due to this load. Following the concept of the first and second load discussed above we shall do the following: 1. Remove the load w and apply a concentrated unit load at C in the direction of the desired deformation (i.e. horizontal). 2. Now apply the second load which is w. Due to the load w alone, the deformations of the structure will be identical to those that would result if only w was applied. This is ensured due to the principle of superposition. Thus, at point C the structure will move by an additional movement of . Of course Figure 12: Simple example where the this is an additional movement because the point C application of Virtual Work is applied. has already moved due to the unit load. Let us consider now the work which is similar to of Figure 11. This is the work performed on the first load (i.e the unit load in our case) due to the deformation caused by the second load (i.e w in our case). This work is: . 3. Since for every work we consider we must have equivalency between the external work and the internal work , we must now consider the internal work that corresponds to of step 2. Following similar energy definitions as in 4.1 we have: (2)

where , , and are the internal axial force, shear force and bending moment respectively due to the unit load, and , and are the internal axial force, shear force and bending moment respectively due to the actual load of the problem w. The integration occurs over the Advanced Structural Analysis Notes by Professor P. D. Kiousis Last Update: January 21, 2014 Page 12

entire length of the structure L (not just the part that carries the load). As in the case of total work discussed in section 4.1 the axial and shear force contributions are typically very small for beams and frames (but not for trusses where the axial force term is the dominant one). Thus, the above equation is typically expressed as: (3)

4. Given that external and internal work are equal, i.e.

, we have: (4)

5. The method described above is typically called the Virtual Work method since the work we examine is not real, given that the unit load is not real. It is also called the Unit Load method for obvious reasons. 6. Generically, for a given a structure that carries a load system S, for which we would like to calculate the deflection at point A in a specific orientation (we can call this degree of freedom or DOF 1 - very convenient for systems of multiple redundancies), we follow these steps: a. Load the structure ONLY with a unit load applied at point A in the orientation of the deflection to calculate. This means that if the desired deflection is rotation, the unit load is a unit moment. Evaluate the bending moment diagram of the structure m. b. Load the structure ONLY with the actual load (load system S). Evaluate the bending moment deformation M. c. Calculate the deflection by using Equation (4).

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 13

5. Deflections Using the Virtual Work


5.1 Deflections of Beams and Frames due to external loads
There exist many techniques to calculate deflections of beams and frames. One of the most convenient approaches is the Virtual Work, also known as the unit dummy load method, which as stated earlier is based on Equation (4). Typically, is constant within a structural element and as a result expression (4) and be rewritten as: (5)

In this case, the integral Table A1.

can be obtained from the tabulated values presented in Appendix

Example 1
Calculate the vertical deflection at point C of the beam shown in Figure 13. 1. Remove the uniformly distributed load w and apply a unit load at C in the vertical direction. The bending moment diagram consists of two linear sections with the following values: The m diagram is presented in Fig. 13D. 2. Develop the bending moment diagram M for the distributed load. This consists of two 2nd order parabolas as demonstrated in Fig. 13B. As in the case of the unit load, and , due to boundary condition restrictions. The moment at B is calculated as the sum of all moments about B of all forces AFTER B: . 3. Depending on the way we want to use Table 1 for integration, we may wish to know the moment at midspan of AB: 4. Based on Equation 5, we can calculate the vertical deflection at C as follows: Figure 13: Example of deflection calculations using Virtual Work.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 14

Based on Table 1, Or

( (

)(

)(

Alternatively, we note that the moment diagram in AB is the superposition of a linear moment diagram and a quadratic and as a result, the deflection can be calculated as: + +

Based on Table 1:

Substituting term, we get: or

)(

)(

)(

))

Example 2
Consider the frame of Figure 16. It is loaded with a vertical load P at point C. We would like to calculate the rotation at point C. The moment of inertia of members AB and BC is respectively IAB and IBC. The dashed lines represent the bottom fibers at AB and BC so help define the proper signs of the moment diagrams.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 15

To evaluate the rotation at C we follow these steps: 1. Find the bending moment diagram M due to the actual load (i.e. P). The M diagram for each member is presented to the right of the structure with the load P in Figure 16. 2. Apply a unit moment at point C (i.e. a unit load at point C in the direction of the desired deformation). The corresponding moment diagram m for each member is presented to the right of the structure with the unit moment in Figure 14. The rotation at C can be calculated as: Based on Table A1 of the appendix: ( )( ) ( )( )

Figure 14: Example of deflection calculations using Virtual Work.

Note from the above calculation that when L =H the column contribution to the rotation is twice as large as the beam contribution. Thus, if we wish to reduce the angle NC, we can achieve it more efficiently by increasing the rigidity EI of the column rather than that of the beam. Following a similar process, if we want to calculate the horizontal deformation at point C, we would apply a horizontal unit load at point C rather than a unit bending moment. The result of this calculation would lead to the conclusion that the horizontal deformation of point C depends entirely on the column irrespective of the relative size of H and L.

5.2 Deflections of Beams and Frames due to internal discontinuities and temperature effects.
Internal discontinuities and temperature effects also cause displacements within a structure. However, if the structures are statically determinate, these displacements develop without the development of internal forces. Examples of such internal discontinuities and the resulting structural movements are presented in Figure 15. In the first example, joint D was constructed by joining AC and CD with an angle of . In the second example, the long member CD was joined with a shear offset . Finally, in the third example, member AC is longer than designed by , either as a result of a construction defect, or because on a temperature increase. Advanced Structural Analysis Notes by Professor P. D. Kiousis Last Update: January 21, 2014 Page 16

Figure 15: Example of deflections due to internal discontinuities and temperatures.

All cases result in rotation of the structure and displacement of nodes. In all three cases, we would like to know the horizontal moment of node B, .

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 17

Given that that internal discontinuities of statically determinate structures result in no internal force, on may wonder how can equation (4) be applied, since in this example, the actual load is the discontinuity, which results in ZERO magnitudes for , , and . This is issue is easily resolved, if we go back to the actual meaning of the principle of virtual work rather than its specific mathematical implementation: We examine the work performed by the deformations of the actual loads on the internal and external forces of the unit load. To calculate the horizontal displacement of point B, we remove the actual load and apply a unit load in the horizontal direction at point B. The solution of this problem, along with the diagrams of internal axial forces , shear forces , and bending moments , is presented in Figure 16. Let us consider now the effects of the angle defect (increased joint angle by ) at point C. Notice that the structure continues to be underformed (Figure 15), with the exception to point C. Thus, the only internal work that is performed by the structural deformations to the internal forces of the unit load is equal to the work of (the internal moment at C due to the unit load) as it rotates by . Note that is in the direction of positive moment. As a result, it was given a positive sign to secure proper sign for the internal work, which is now calculated as The external work, of course is It is concluded: . Figure 16: Axial, Shear, and Bending Moment Diagrams of the frame due to unit load at B.

Next, let us consider the effects of the shear offset at midspan of BC. Similarly to the previous example, the structure remains undeformed at all points but the offset point, which has a shear deformation of . In this case, the internal work is equal to the internal shear force at this point times the shear deformation at this point: ( )( ) . Note that the shear offset is in the direction of positive shear force, so it was given a positive sign to provide the proper sign of the work. The external work, of course is . It is concluded:

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 18

Finally, let us consider the effects of the elongation of member AC. Similarly to the previous example, the structure remains undeformed at all points but over the length AC, which is elongated by . In this case, the internal work is equal to the internal normal force times the member elongation: ( )( ) . Note that the axial elongation is in the positive direction of axial force, so it was given a positive sign to provide the proper sign of the work. The external work, of course is . It is concluded:

5.3 Deflections of Non-Prismatic Beams


Consider the non-prismatic beam shown in Figure 17. The cross-section is rectangular of constant width b, but varying depth h. For this specific example, we have a linear variation of depth as a function of position x. Given that ( ) , the moment of inertia of the cross-sections varies along the length of the beam as: ( ) ( )

To calculate the vertical deflection at the right end, we follow the typical steps described earlier. The vertical deflection is calculated as:
( ) ( ) ( )

Figure 17: Non-prismatic beam

For prismatic beams, where the moment of inertia is constant along the entire length of the beam, the above expression simplifies into:
( )

( ) ( )

, and the calculation of this integration can

be aided by tabulated expressions such as the ones presented in Table A1 of the Appendix. In the example presented here, the moment of inertia ( )in the denominator is not constant, resulting in a harder computation. For the specific example, we have ( ) and ( ) . Thus, the expression for the vertical deflection of the left point becomes:

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 19

Whereas analytical solutions of such integrals are often possible, it is more common to rely on computational approaches to achieve a solution. A simple numerical Integration method is summarized in the Appendix. Let us now examine our problem in more detail. To achieve a numerical solution we need to have specific numbers for our calculations. For example: m, m; m; m, , which results in ( ) ; and kN/m. The modulus of elasticity of the material is GPa. Note that the depth of the cross-section is now precisely known as a function of x: ( ) , and thus the deflection at the free end becomes:

The exact integration of this expression results in . The numerical integration discussed in the appendix results in m when the length is divided in 80 intervals, and it results in m, when the length is divided in 8000 intervals. We can see that non-prismatic bars can be an efficient way of controlling deflections (if done properly). In our case, by increasing the moment of inertia from left to right we followed the trend of the bending moment diagram (increasing the moment of inertia , with increasing bending moment ) which resulted in favorable deflection calculations.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 20

6. Statically Indeterminate Structures- Virtual Work and Superposition


6.1 Introduction
Statically indeterminate elastic structures can be solved using the principle of superposition and equations that express the compatibility of deformations. In such an approach of solving statically indeterminate structures, redundant reactions or internal forces are identified and their corresponding deformations are released for the purpose of producing statically determinate models with unknown loads that can be evaluated so as to produce compatible deformations.

6.2 Beams One Degree of Redundancy


We shall develop this subject using multiple examples. Example 1: We shall start with one of the simplest statically indeterminate structures (Figure 18). Beam AB is statically indeterminate with degree of redundancy equal to 1. This is easy to identify, since there are four potential reactions ( , , , and ), and only three equations of equilibrium (e.g. , and .). To solve this problem, we must first identify the redundant force. A large number of possibilities exist. The redundant force (i.e. a restriction that is not necessary to make the structure stable) can be any of the reactions other than , or any internal restriction, such as the moment M at any point along the length AB. It is quite common for problems of this type to select the redundant force as the reaction . Thus, our problem as described in Figure 18A is equivalent to the statically determinate structure shown in Figure 18B, where the support at A has been replaced by its unknown reaction. The principle of superposition allows us to analyze our structure of Figure 18B as the sum of its loading Figure 18: Pinned-Fixed Beam constituents, that is the distributed load w (Figure 18C) and the reaction (Figure 18D). Similarly, the cantilever structure loaded by the reaction force is analyzed as a structure loaded at point A by a unit load with a solution that is then multiplied by . The cantilever structure loaded by the uniform load deflects at point A by an amount . Similarly, point A deflects due to the unit load at A by an amount . Thus, the deflection at A due to is equal to . If we assign proper signs to the deflections, typically positive in the direction of the unit load, then we can produce the equation of compatibility of deformations as follows:

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 21

Note that the above equation of compatibility simply states that since the sum of problems of Figures 18C and 18D is equivalent to the problem of Figure 18A, then the deformation at point A = must be equal to zero, given that point A rests on a roller and does not move vertically. The solution of this problem has been conceptually completed at this point, since the quantities and are easy to calculate for statically determinate structures. Using the concepts of Chapter 5 on virtual work, a) we produce the bending moment diagrams of the cantilever beam subjected to the distributed load w and the unit load (Figure 19) and b) we calculate the deflections and : ( ) Thus, Similarly, ( ( )) Thus, We can now calculated By calculating the reaction , the problem is now a statically determinate cantilever structure and an end concentrated force . The shear force ( ) Figure 19: M and m diagrams for the Cantilever Beam ( ( ) ) ( )

subjected to a uniformly distributed load

and bending moment diagrams are presented in Figure 20. It is interesting to note that the solution of this beam is not dependent on the magnitude of EI, even though we used deformations in producing it.

Figure 20: V and M diagrams for the pinned-Fixed Beam

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 22

Example 2: For the next example, we shall consider the same problem, other than the fact that the beam now consists of two beams, with different cross-sections, as demonstrated in Figure 21A. As discussed earlier, the beam is statically indeterminate with degree of redundancy equal to 1. We shall follow the identical process to solve this problem as we did in the previous example. This process includes the analysis of the structure of Figure 21A into its equivalent statically determinant structure of Figure 21B, and then its components in Figures 21C and 21D. The process leads to the identical equation of the compatibility of deformations

The only difference between this problem and the previous one comes on the evaluation of and . Our integrals are now slightly more complicated because EI is not constant over the entire length L.

Figure 21: Pinned-Fixed Non-Prismatic Beam

The bending moment diagrams M of Figure 21C and m of Figure 21D are presented in Figure 19. To calculate the deflection Thus, , we do the following: ( ) ( ) ( ) ( ) ( ) ( )

)(

)(

) [(

)(

) ]

Or

Similarly we calculate the deflection

as follows:

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 23

Thus,

[ ( )]

[ ( )]

[ ( )]

)(

( ) ( ) [(

Or Thus,

Note that the magnitude of the reaction

is smaller than that of Example 1 (

), where EI was

constant over the entire length of the beam. We can now calculate the bending moment value at point B: ( ) .

Note, again that the moment at B is larger in magnitude than that of Example 1 ( ). This is not surprising once we found that the reaction at A was smaller. The reaction at A provides the positive contribution in addition to the negative cantilever value of . The smaller the reaction at A, the more negative the moment at B becomes. The comparison of the moment diagrams between Examples 1 and 2 is presented in Figure 22. The observed behavior is an example of a well-known principle of structural mechanics: The internal forces of a statically indeterminate structure tend to go where the stiffness is. The part of the structure close to the support is stiffer. Thus it has attracted more shear force and more bending moment. The implications of this behavior are very important and we see them in structural decisions of what structure to solve all around us.

Figure 22: Effects in moment diagram due to relative stiffness change.

6.3 Lessons from the non-Prismatic Beam Outcome


The statement The internal forces of a statically indeterminate structure tend to go where the stiffness is has significant implication on how we design structures. Advanced Structural Analysis Notes by Professor P. D. Kiousis Last Update: January 21, 2014 Page 24

Let us revisit the solution that we achieved for the prismatic pinned-fixed beam, which is demonstrated in Figure 22. The span moment is , while the fixed end moment is . Thus, our entire prismatic beam must be designed to resist the maximum moment, which occurs only at one point. This may prove to be wasteful and expensive, especially in long structures. One may conclude that what we should pursue is a solution where we decrease the moment at the fixed end, while we increase the moment at the mid-span until they are equal. This is not difficult to achieve mathematically. The required support reaction must be such that the span moment is equal to the negative of the fixed end moment .

This is a quadratic equation in

. It has a positive and a negative root. The negative root is rejected as ) , where . is at 0.4142 L from A. to achieve this value?

meaningless for our problem. The positive root is ( The resulting moments are

The next question that one must address is how do we force the reaction

An easy (and by no means the ONLY) answer is to use the concept that the moment goes where the stiffness is. Thus, following an approach opposite in concept to that of the example 2, where the beam has a smaller moment of inertia close to the fixed end, may give us the solution that we need. This indeed works. Consider the example of Figure 23, where the stiffness of the beam close to the support has a moment of inertia that is a fraction of the moment of inertial at the midspan: . If , then we can calculate that forces the reaction to be , and thus have equal span and support moments.

Figure 23: Non-prismatic beam, weak at the fixed support.

HOWEVER, is this desirable? Clearly the support is now weaker than the mid-span (smaller ). What is therefore the benefit of having the same moment at the support and at midspan? Clearly this is not a good solution. A better solution would be one that results in moments at the span and at the support that are proportional to their moments of inertia. It would be even better if these moments were proportional to the section moduli. This approach, conceptually, is described in Figure 24. Note however, that even though we have achieved efficiency in the sense that our moment diagram is more consistent with the stiffness and strength distribution of our beam, this is not necessarily a good solution, because it is achieved by constructing most of our beam using a big (and thus expensive) section. Thus, an alternative way of thinking is desirable. Advanced Structural Analysis Notes by Professor P. D. Kiousis Last Update: January 21, 2014 Page 25

We prefer one of the following two options: 1. We achieve moment balance, where the positive and negative moments have the same magnitude, while we maintain a prismatic shape (same strength for same moment). It is easy to force our beam behave as such, by placing a pin at , which is the location of zero moment when . 2. We exaggerate the negative moments, while reducing the positive moment by using large stiffness close to the support. Now, we achieve conceptually moments in proportion to the placement of stiffness, while at the same time we construct only a small part of our beam using the strong and expensive cross-section. These two approaches are presented in Figure 24. The use of the pin allows a design with the minimum bending moment diagram over the entire beam length, which is best accommodated by a prismatic beam design, i.e. a constant cross-section designed to carry a moment of . The non-prismatic approach of Figure 24 is also very efficient because it allows 80% of the length of the beam to be designed based on a very small moment ( ), while the 20% of its length must accommodate the much larger bending moment of . Figure 24: Efficient alternative designs of the pinned-fixed beam.

6.4 Frames One degree of Redundancy


Let us consider the portal frame of Figure 25, which is statically indeterminate with a degree of redundancy equal to 1. We can solve this problem using the same principles as in section 6.2 for the statically indeterminate beam. We select the horizontal reaction at B to be the redundant force, thus producing the superposition of two statically determinate problems, as demonstrated in Figure 25. Since our system has released the horizontal deformation at point B, we must enforce zero horizontal deformations, resulting in the following equation of compatibility of deformations:

where, Based on Figure 25,

and

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 26

Similarly, or ( )

Thus,

Figure 25: Portal Frame with one degree of redundancy. The resulting bending moment diagram is presented in Figure 26. The following are interesting cases to examine: Case 1: (Very small girder stiffness

compared to the column stiffness). In this case, becomes:

and thus the end moment on the girder becomes , which is the Figure 26: M diagram of pinned portal frame bending moment at the support of a beam with both ends fixed. Case 2: (Very large girder stiffness compared to the column stiffness).

In this case, , and thus the end moment of the girder becomes zero, making the girder CD a simply supported beam. These findings are not arbitrary. Instead, they express the natural behavior of this frame. Advanced Structural Analysis Notes by Professor P. D. Kiousis Last Update: January 21, 2014 Page 27

In the first case, where the columns are much stiffer than the girder, the load on the girder causes bending deformations on the girder. However, the very compliant girder cannot rotation the stiff column at points C and D. As a result, the girder CD behaves as if it is fixed at both ends. In the second case, where the girder is much stiffer than the columns, the opposite process occurs. The load causes bending deformations on the girder including rotations at the ends C and D. However, these rotations are very small for the very compliant columns to register any appreciable reaction. Thus, no appreciable moments develops at the ends of the girder CD, which thus behaves as a simply supported structure, while the columns are loaded only axially. We can conclude that as the columns start gaining stiffness compared to the girder, the end moments of the girder increase, and eventually tend to the value of full fixation when the columns become infinitely stiff compared to the girders. Relatively small stiffness columns are commonly selected in structural systems of low physical height, ensuring small moments in the columns and their connections to the girders. At the same time, braces for steel structures and shear walls for concrete structures are used at select places to take most of the shear loads caused by large seismic and/or wind loads.

7. Statically Indeterminate Structures Deformation Method


7.1 Development of the Concept.
Let us consider the two-span continuous beam of Figure 27a. Let us also assume that we fix the rotation of the mid-support (point B). In this case, when the distributed load is applied, beams AB and BC behave as pinned-fixed beams. Thus, the moment of point B for beam AB is equal to to (clockwise) and the moment of point B for beam BC is equal

(counter clockwise). These two moments are unbalanced in the general case, and the

difference is taken by the rotational fixity device that enforces zero rotation at B. If this fixity is removed, node B will naturally rotate in the direction of the larger moment (in this example clockwise) and the moments will balance. We observe that each of the sections AB and BC of this beam behave as if they are pinned at one end (A or C), fixed at the other (B), and subjected to two loads: a) the uniformly distributed load , and a rotation of the fixed node B. We can take advantage of this observation as follows: a) We assume that the rotation at B is clockwise. b) We use the convention that any moment on the beam that is clockwise is positive. c) We evaluate the moment at each beam end as a sum of the moments due to the load and due to the rotation. Advanced Structural Analysis Notes by Professor P. D. Kiousis Last Update: January 21, 2014 Page 28

d) Since the same moments acting on the beams, act also on the node, they must satisfy ). equilibrium ( e) The equations of equilibrium (one for each internal node) result in a number of equations that is equal to the number of unknown rotations. The solution of these equations results in the solution of the unknown angles of rotation. We implement the above steps on the beam of Figure 27. The convention for the beam moments is as follows: is the moment at node that belongs to the beam . Based on the table of section A3 we have: 1. 2. 3. Equilibrium requires that: . Thus: ( )

4. The above equation allows a solution for 5.


( )

can now be substituted in steps 1 and 2 to calculate and .

Figure 27: Development of the concept of the deformation method Continuous beam example.

7.2 Beam Example


We shall implement the deformation method to solve for the end moments of the continuous beam of Figure 28. The beam consists of three spans, as analyzed in Figure 28b, where each span is subjected to its actual load plus fixed end rotations. The following steps are followed to produce the solution: 1. Moment equations at each beam end ;

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 29

2. Nodal Equilibrium

or where and Figure 28: Application of the deformation method Continuous beam example. or 7-2-Ex2 7-2-Ex1

3. Solution of the System of Equations Solution of the system of equations (7-2-Ex1) and (7-2-Ex2) results in and .

4. Evaluation of the end moments Back substitution of the above values for and into the equations for end moments result in: ; ;

5. Assign proper sign to the moments. Note that the above moments have been produced based on the sign convention that clockwise moments are positive, and are shown with their proper orientation in Figure 29. Clearly, all moments

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 30

operate to put the bottom fiber of their corresponding beam in compression. Thus, they are all negative:

6. Evaluation of shears

) Figure 29: Calculated support moments with their proper orientations.

( (

) )

( (

) )

7. Evaluation of Max span moments @ @ @ 8. Bending moment and shear force diagrams The bending moment and shear force diagrams are presented in Figure 30 from point A. from point B. from point C.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 31

Figure 30: Bending moment and shear force diagrams of continuous beam

7.3 Frame Example


Consider the frame of Figure 31. We shall begin our discussion with the engineering decisions that were made to define this problem. a. The load source is the reaction of a beam that cannot be supported at this point by a column. b. The types of supports are decided based the structural material (concrete or steel), the foundation conditions, and the effects that a fixed or pinned support have to the overall behavior of the structure. Foundations are typically made of concrete. A concrete column is constructed fixed to the foundation in most cases (although not always) as illustrated in Figure 31a. A steel column can Advanced Structural Analysis Notes by Professor P. D. Kiousis

Figure 31: Bending moment and shear force diagrams of continuous beam Last Update: January 21, 2014 Page 32

be constructed fixed or pinned to the concrete foundation (Figure 31b). If the column flanges are welded (welds to develop moment) then the connection is fixed. If the column flanges are not welded then the connection is pinned. However, the fact that the column to footing connection is fixed does not guarantee that the support is fixed. Support fixation requires, in addition to column-to-connection-fixation, that the footing is wide enough and the foundation soil is stiff enough to develop the fixation moment without appreciable rotation. c. The relative stiffness of the girder to the column is designed to improve design efficiency. That is, produce a bending moment diagram, where the developed bending moments are efficiently matched with the element moment capacities. Making a stiff column compared to the girder (e.g. ) results in a node B rotation fixation, and as a result a moment at B equal to and a moment at D equal to . We conclude that the moment at B is the dominant moment in the girder. It is also the dominant moment of the column. Thus, the moment demand of both the girder and column is the same, BUT, the column is must stronger than the girder, and as a result, this is an inefficient design. We selected a ratio =3. Making the girder stiffer results in a smaller moment at B and a larger moment at D. This is an efficient approach, because the stronger girder is designed by the larger moment at D and the weaker column is designed by the smaller moment at B. In addition, producing a design where the moment at B is small has the additional advantages of having a small moment at the connection (the weak link of most structures), and a small overall moment to the column (helps resisting buckling). In deciding the method of solution, we conclude that this problem is a good candidate for a displacement method solution given that the only unknown nodal deformation is the rotation , while its degree of redundancy is 2. The solution process is as follows: Figure 31: Examples of column to footing connections

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 33

1. Moment equations at beam ends

( )

2. Nodal equilibrium

3. Solution for 4. Evaluation of the moments

5. Assign proper sign to the moments

6. Evaluation of Shears ( ( ( ( ) ) ) )

7. Evaluation of moment at D

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 34

8. Bending moment and shear force diagrams The bending moment and shear force diagrams are presented in Figure 32.

Figure 32: Bending moment and shear force diagrams of continuous beam

7.4 Second Frame Example


Let us reexamine the frame of the previous example, with one change: The pin at support C is now a roller. The selection of supports and relative girder-tocolumn stiffness is selected based on the logic discussed earlier. The problem has now been reduced to one degree of redundancy, and is easier to solve using the force method. On the other hand, its degree of difficulty has been increased when considering the displacement method. In the previous example, there was only one unknown deformation, the rotation . The release of horizontal restriction at point C has resulted in an additional unknown deformation: the horizontal displacement of the girder BC. As a result, this specific problem is easier to solve using the

Figure 33: Bending moment and shear force diagrams of frame

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 35

force method. However, it is solved here using the displacement method to demonstrate how to use this method when unknown displacements exist. The deformations of each element are as shown in Figure 34. It can be seen that the column AB is subjected to a top node rotation and displacement. The bending moment and shear force at node B are as follows:

( )

Notice that, as opposed to what has been done in the previous examples, our calculations in this problem include shear forces. Since our unknown deformations include the rotation and the horizontal displacement . The equilibrium of moments at point B produces one equation in terms of and .

Figure 34: Deformations, bending moments and shear forces developed at the ends of each frame element

A second equation is needed, which cannot be a moment equilibrium expression (there are no other left). Instead, we note that there are two shear forces developed at the top of the column AB: The shear due to: ( ) and the shear due to ( ). These two shear forces are also applied with equal magnitude, but opposite direction, at point B, at the bottom of beam BC. Equilibrium of forces for this beam in the horizontal direction demands that these shear forces add up to zero. Thus, the equilibrium equations to solve this problem are:

The solution of the above system of equations results in: Last Update: January 21, 2014 Page 36

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Substitution of and into the moment and shear equations above results in:

Applying the bending moment diagram sign convention results in:

The shears can now be calculated:

The bending moment at D is

The shear force and bending moment diagrams are shown in figure 35.

Figure 35: Bending moment and shear force diagrams of frame

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 37

8. Trusses An Alternative Approach


8.1 A Quick Implementation of the Method of Joints
Trusses are commonly solved with the method of joints. This is an approach that simply examines equilibrium one joint at a time. Because a joint has not geometry, its equilibrium is satisfied by only two equations ( ). As a result, efficient implementation of this concept requires that we always have a node, where only two unknown forces converge. In most practical applications (i.e. for commonly used trusses), the implementation of this method can be done implicitly as will be demonstrated in the example that follows.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 38

A. APPENDIX
A.1 Integration Table

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 39

A.2 Numerical Integration


A.2.1 On point compound Integration

This is one of the simplest compound integration techniques to calculate a definite integral. In the example of Figure A1, a function ( ) is to be integrated from points to . To calculate this integration, the domain [ ] is divided in a number of equal intervals. In this example, we have selected 13 equal intervals, introducing the new points . For each point, we evaluate the function: , and we approximate our integral as Figure A1: One-Point Compound Integration. the area of all the shaded rectangulars, as shown in Figure A1. Considering that the points are equidistant, the base of each rectangular is

Thus:

( )

, where =number of points.

Clearly, the larger the number of intervals that we divide our domain, the closer our equation comes to the theoretical definition of the definite integral, and the more accurate our results. This is demonstrated in Figure A2, where the equation for the deflection
( )

, of the end point of the

prismatic beam of Section 5.3, is evaluated for the following parameters L=8m; b=0.4m; ho=0.5m; =1/8; w=200 kN/m; and E=20 GPa. The integral was evaluated for different resolutions (magnitude of h) as shown in Table A2. Table A2 h 1 0.1 0.01 0.001 Figure A2: Convergence of One-Point Compound Integration.

0.068653 0.07865 0.079672 0.079774

It is clearly demonstrated that this computational approach converges to a calculated deflection of approximately 0.08 m as the interval of the numerical integration approaches 0. Other compound integration techniques, such as the Trapezoidal rule or Simpsons rule can be used instead of the one point compound rule that was presented here.

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 40

A.3 Deformation Method Primer

Advanced Structural Analysis Notes by Professor P. D. Kiousis

Last Update: January 21, 2014 Page 41

Potrebbero piacerti anche