Download as pdf or txt
Download as pdf or txt
You are on page 1of 48

ANALYSIS OF PLANAR TRUSS STRUCTURES 3-1

CHAPTER 3
ANALYSIS OF PLANAR TRUSS STRUCTURES

A simple truss is generated by successively attaching two members connected by


one joint to a basic three-member, three-joint, triangular unit. They are lightweight and
also rigid structures meaning they do not freely collapse when load is applied. A
compound truss is a multiplex of simple trusses that is also rigid. Bridge trusses, roof
trusses, cantilever trusses, and other types of trusses are common structures. Their
internal forces are identified here as bar forces and they are the main design objective for
truss structures.
Section 3-1. A Review - Characteristics of Idealized Truss Structures
Idealized truss structures are characterized by:
(1) Two-force members (straight, slender), Unknown bar Available
(2) Frictionless pin joints, forces plus equations of
(3) Loads applied only at joints. reactions equilibrium
(m+r) (2n)

21 bar forces 12 jo int s


+3 reactions ×2 eqn′s / jo int
24 unknowns 24 equations

24 bar forces
+4 bar forces 24 equations
28 unknowns

24 bar forces
−1 bar forces 24 equations
23 unknowns

24 unknowns 24 eqn's (not


all independent)
Idealized examples of a bridge truss structure
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-2

Static determinacy implies a rigid structure with no redundancy as follows:


(1) The number of unknown bar forces (m) and unknown reaction forces (r = 3) is
equal to the number of equilibrium equations (2n) (twice the number of joints n)
available to solve for them, i.e., m+3 = 2n,
(2) The 2n equations of equilibrium are linearly independent.
For example, there are 24 unknowns and 24 equations of equilibrium in the statically
determinate bridge truss shown. When one member fails, the structure becomes non-
rigid, meaning the structure is immediately vulnerable to collapse.
Static indeterminacy implies that the number of unknown bar forces and reactions
is greater than the number of equilibrium equations available to solve for them, i.e.,
m+3 > 2n; e.g., 28 versus 24 in the example indeterminate bridge truss shown.
Moreover, this structure has static redundancy of order 4. Structural redundancy is a
safety feature, and an objective of standard practice in engineering analysis and design.1
When solving a statically determinate truss structure by hand it is usually
advisable, but not always, to calculate the reactions first using a FBD of the entire
structure. Thus, for example, the reactions at the left and right supports are R = 5P 2
acting upwards on the bridge truss. The two methods, which are developed in elementary
engineering mechanics, that are used to solve for the bar forces are the method of joints
and the method of sections. Either or both of these methods are generally applied,
initially making use of support reaction data. These two solution methods are also
important for hand-checking finite element solutions. They are briefly reviewed in the
following for the statically determinate bridge truss shown above.
A bar force or internal axial force is denoted F. It is a stress resultant or integral
of the axial stress distribution acting over a normal section through the bar, and it is a
constant for two-force members.
In the FBD's shown below, all unknown bar forces are assumed to be in tension to
promote an orderly solution procedure. For example, the subsequent solution confirms
bar BC is in tension so its bar force is in the direction shown at joint C. Bar BD is found
to be in compression so its bar force acts opposite to the direction shown at joint B.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-3

∑F V
= 0 ⇒ FBC = P ∑M E
= 0 ⇒ FBD = −4Pl h

∑F H
= 0 ⇒ FCE = FAC
∑F V (
= 0 ⇒ FBE = 3 h + l
2 2
)
1 2
P ( 2h )
∑F H
= 0 ⇒ FCE = 5 Pl ( 2h )
Method Method
of of
joints sections

Solution methods for truss structures

The solution FBC = P is obtained from a FBD of joint C by applying the

condition of vertical equilibrium, one of the two equilibrium conditions available for the
FBD’s concurrent force system. The other condition yields FCE = FAC . The solution

FBD = − 4Pl h is obtained from a FBD of the truss to the right of a vertical section

through bars BD, BE and CE. The condition of moment equilibrium is strategically
applied with respect to joint E to eliminate the other two unknown bar forces FBE and FCE
from the resulting equilibrium equation. This is one of three equilibrium conditions
available for the FBD's planar force system. The remaining two conditions are used to
solve for FBE and FCE.
These elementary solution methods are further reviewed in the next two
examples. The second example extends the review and considers the stability of a
member’s force system. It’s important to label bar forces either tension or compression.
===============================================================
Example 3-1.
(a) Find the bar forces in members BC, CD and DE of the cantilever truss
structure shown.
A 12 ' B

9'
C
E 12 ' D 12 '
6 kips
9 kips

1
The Federal Highway Administration defines a fracture critical bridge as one having a steel tension
member whose failure would cause a portion of or the entire bridge to collapse. Such a bridge generally
lacks sufficient static redundancy.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-4

FBD of joint C (a force polygon can also be used):


3
∑ Fy = 0 ⇒ 5 FBC − 6 = 0 ⇒ FBC = 10 kip T
4
∑ Fx = 0 ⇒ − FCD − 5 (10 ) = 0 ⇒ FCD = −8 or 8 kip C

FBD of section through AB, AD and DE:


∑ MA = 0
9 × 12 + 6 × 24 + FDE × 9 = 0
FDE = −28 or 28 kip C

Observe that the structure’s reactions


were not needed for these calculations.
(b) Determine the cross-sectional areas
for members BC, CD and DE when the
design stress is 20 ksi in tension
members and 5 ksi in compression members.

10 8 28
ABC = = 0.5 in 2 ACD = = 1.6 in 2 ADE = = 5.6 in 2
20 5 5

These solution methods tacitly assume that force systems acting on FBDs are in
stable equilibrium. However, e.g., the top chords of the bridge truss, such as member
BD, or the bottom chords of the cantilever truss, such as member DE, are generally high
compression members. And when compression members are slender, they are subject to
instability and potential failure by elastic buckling. Their force systems may or may not
be in stable equilibrium.
Slender compression members may be checked against elastic buckling using the
classical Euler buckling load formula for an axially loaded simply supported member,
Pcr = π 2 EI l 2 (critical load)
Here, EI and l are, respectively, the bending rigidity with respect to the minimum
moment of inertia axis of the member's cross-section, and the (unbraced) length of the
member. The check for an arbitrary slender compression member (e) is,

e ⎪ < Pcr ⇒ stable equilibrium, no elastic buckling failure
F( ) ⎨
⎪⎩ > Pcr ⇒ unstable equilibrium, elastic buckling failure
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-5

The total potential energy of a slender


compression member is written as a function of axial
displacement u (i.e., axial shortening of the member in
this case). According to the principle of minimum
potential energy, when the first derivative with respect
to u of the member's total potential energy Π ( e) ( u ) is

zero, i.e., ∂Π ( ) ∂u = 0 , the member’s force system is in


e
Stable and unstable equilibrium
equilibrium. However, the type of equilibrium is not
indicated by this stationary condition. To determine the type of equilibrium, one must
also consider the second derivative as illustrated in the sketch. According to the

principle, when the second derivative is positive, i.e., ∂ 2Π ( ) ∂u 2 > 0 , the member's force
e

system is in stable equilibrium, and conversely when it is negative, i.e., ∂ 2Π ( ) ∂u 2 < 0 ,


e

the member's force system is in unstable equilibrium (Section 2-2.1).2

===============================================================
Example 3-2. Slender truss members subjected to compression forces are vulnerable to
elastic buckling and must be checked. This applies to members CD and DE of Example
3-1. Let the truss members be fabricated of aluminum alloy 6061-T6 with Young's
modulus E = 10 × 10 3 ksi .
Assume that the members are solid circular bars so that any diametrical axis of
their cross-section is the minimum axis. The tabulated results of the check are:

Member F(e) A 2r I = πr4 4 l Pcr = π 2 EI l 2 Safe?


(kip) (in2) (in) (in4) (in) (kip)
CD -8 1.6 1.427 0.2037 144 0.969 No
DE -28 5.6 2.67 2.495 144 11.875 No

The last column in the table indicates that, according to the minimum potential
energy principle, the equilibrium of these members is unstable (i.e., F ( e) > Pcr ), and failure
due to elastic buckling will occur. Therefore, when the members are in fact the
prescribed circular aluminum bars, the tacit assumption that all members are in stable
equilibrium made in Example 3-1 would be incorrect.
2
An advantage of the principle of minimum potential energy for structural analysis is that it addresses
directly not only equilibrium, but also the type of equilibrium, distinguishing between stable and unstable
equilibrium, as discussed for bar BD. Stability of equilibrium is not addressed directly by either the
principles of equilibrium or virtual work. Nonetheless, the latter principle remains a powerful alternative,
fundamental basis for computational structural analysis and FEM. It is based on the fundamental theorem
of variational calculus as discussed in Chapter 2. See, also for example, K. D. Hjelmstad, Fundamentals of
Structural Mechanics, 2nd Ed., Springer, New York, 2005.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-6

If instead the members are standard aluminum I-beam cross-sections (of


reasonably comparable weight), for which I y is the minimum moment of inertia, the
results of the check are:

Member F(e) A* Ix * Iy * l Pcr = π 2 EI y l 2 Safe?


(kip) (in2) (in4) (in4) (in)
(kip)
CD -8 3.5 (13.94) 2.29 144 10.90 Yes
DE -28 5.256 (59.69) 7.30 144 34.75 Yes
* Properties of Aluminum Association Standard I-beams.

Thus, when using these aluminum I-beam shapes the compression members in the truss
of Example 3-1 are in stable equilibrium (i.e., F ( e) < Pcr ), and the static equilibrium
analysis of the truss is valid.

In general, unless otherwise specified, it is assumed going forward that all


structures are stable, meaning that their force system is in stable equilibrium.
Alternatively, it is assumed that the second derivative of the total potential energy for
these structures is positive (Section 1-5).
Joint deflections and bar forces in truss structures may be determined using
alternative procedures based on the principle of virtual work, e.g., Castigliano's 1st
theorem.3 These and other examples of classical analytical and numerical methods are
presented in textbooks on matrix structural analysis for idealized discrete systems,
including truss structures.4
The present goal is to provide a generally applicable numerical solution
experience with FEM technology and high-end commercial/industrial FEM software.
FEM is indeed the method of choice for modern structural analysis and design in the
workplace.
Finally, as indicated, structures are generally idealized to be statically
indeterminate in practice, and it is worthy of note that FEM does not distinguish between
statically determinate and statically indeterminate idealizations. This is an advantage that
FEM has over classical solution methods. The reason for this has been discussed (see
Section 1-2.2.3). The gist of it being that irrespective of whether the structural

3
J. S. Przemieniecki (1968). Theory of Matrix Structural Analysis, McGraw-Hill Book Company, New
York, 468 pages, 1968.
4
L. P. Felton and R. B. Nelson (1997). Matrix Structural Analysis, John Wiley & Sons, Inc., New York,
700 pages, 1997.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-7

idealization is statically determinate or statically indeterminate a finite element system of


equations KD =Q satisfies: (1) boundary conditions, (2) nodal compatibility conditions,
and (3) nodal equilibrium conditions. Yet, engineers do need to distinguish between
these two structural idealizations so that they may design-in sufficient static redundancy
for safe transfer of loads to supports.
Section 3-2. Member Coordinate Systems and Transformation Matrices.
There are three coordinate systems for truss structures and their members, each with an
associated transformation matrix.5 Forces and displacements for a truss member
embedded in "basic", "local" and "global" coordinate systems are shown below. The
basic system is one-dimensional, the local system is two-dimensional or planar, and the
global system is either two- or three-dimensional (only the former is considered here).
All quantities are counted positive when in the direction of the positive coordinate axes as
shown in the local and global systems. One selects the positive x -axis direction for each
member so that nodes i and j are uniquely identified for each member. The angle β
shown is the angle between the local positive x-axis and the global positive X-axis, and is
measured positive counterclockwise.
The basic member force, denoted F ( e) , is counted positive when tension. The
basic member deflection is the member’s change in length and is denoted Δ, and it is
counted positive when the member elongates. Nodal displacements and member
rotations are assumed so small that changes in geometry of the truss structure may be
safely neglected when developing equilibrium equations for members and for the
structure. That is, equilibrium equations are developed with respect to the undeformed
configuration of the structure.
Basic force and basic displacement variables number one each for they are
associated with the member, not the member’s nodes. Local and global force and
displacement variables number four each for they are rectangular components associated
with the two nodes of the member, i and j.

5
A fourth coordinate system called a joint coordinate system is reserved for subsequent discussion (Section
3-5). It is used when it is necessary to impose nodal displacement boundary conditions that are not
prescribed in global coordinate directions, but rather in skew coordinate directions.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-8

Basic System Local System Global System

Forces

Displacements

Forces and displacements in truss member coordinate systems

3-2.1 Basic-to-local coordinate transformation. Nodal forces riy′ and rjy′ are zero

since truss members are two-force members. Nonetheless, they are retained as
"placeholders" in the order-4 nodal force matrix in the local system r ′( e) so that it
conforms to the order-4 nodal displacement matrix in the local system d′ .
(1) Forces:

rix′ = −F ( e) ⎧ rix′ ⎫ ⎧ ⎫
⎪ ⎪ −1
riy′ = 0 ⎪ riy′ ⎪ ⎪ ⎪⎪
( e) ⎪ 0 ( e) ( e)
⇒ ⎨ ⎬= F ⎨ ⎬ or r ′ = F t
rjx′ = F ( e) ⎪ rjx′ ⎪ ⎪ 1 ⎪
rjy′ = 0 ⎪ rjy′ ⎪ ⎪⎩ 0 ⎪⎭
⎩ ⎭

FBDs of nodes i and j


This transformation is based on FBD's of nodes i and j, as
shown. That is, equilibrium exists between internal force F ( e) and external nodal force
rjx′ or − rix′ , and overall equilibrium of the two-force member, i.e., rix′ = − rjx′ , is also

indicated.
(2) Displacements: By inspection,
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-9

⎧ dix′ ⎫
⎪ ⎪
⎪⎪ diy′ ⎪⎪
Δ = d ′jx − dix′ ⇒ Δ = ⎢⎣ −1 0 1 0 ⎥⎦ ⎨ ⎬ or Δ = t T d′
⎪ d ′jx ⎪
⎪ d ′jy ⎪
⎪⎩ ⎪⎭
3-2.2 Local-to-global coordinate transformation.
(1) Forces: The local-to-global force transformation at node i is,
riX = rix′ cos β − riy′ sin β
riY = rix′ sin β + riy′ cos β

It is noted that squaring both of these equations and adding will yield,
rix′ 2 = riX 2 + riY 2
The force transformation is the same at node j, so the combination of each transformation
in matrix form is,
⎧ riX ⎫ ⎛ cos β − sin β 0 0 ⎞⎧ rix′ ⎫
⎪ ⎪
riY ⎪ ⎜ sin β ⎟⎪ ⎪
riy′ ⎪
⎪ cos β 0 0 ⎟ ⎪⎨
⎨ =⎜ ( e) ( e) ( e)
⎬ or r = T r ′
rjX ⎬ ⎜ 0 0 cos β − sin β ⎟ ⎪ rjx′ ⎪
⎪ ⎪ ⎜ ⎟
⎪ rjY ⎪ ⎝ 0 0 sin β cos β ⎠ ⎪ rjy′ ⎪
⎩ ⎭ ⎩ ⎭
(2) Displacements: By inspection of the respective components at node i for
example, we get the same local-to-global transformation for displacements as for forces,
diX = dix′ cos β − diy′ sin β
diY = dix′ sin β + diy′ cos β

Thus, the nodal transformation matrix T( e) is the same for displacements and forces,

d = T ( )d′
e

3-2.3 Orthogonality of T(e) and Summary. The member transformation matrix


T( e) is orthogonal. That is, its inverse equals its transpose as follows. Since, in general,
−1
⎛ a b⎞ 1 ⎛ d −b⎞
=
⎜⎝ c d ⎟⎠
( )
ad − cb ⎜⎝ −c a ⎟⎠

It follows that,
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-10

−1 T
⎛ cos β − sin β ⎞ ⎛ cos β sin β ⎞ ⎛ cos β − sin β ⎞
⎜⎝ sin β =⎜ =
cos β ⎟⎠ ⎝ − sin β cos β ⎟⎠ ⎜⎝ sin β cos β ⎟⎠

Thus, the matrix,


⎛ cos β − sin β ⎞
⎜⎝ sin β cos β ⎟⎠

is orthogonal. Finally,
−1
⎛ ⎛ ⎞ ⎛ ⎛ ⎞
T
⎛ ⎛ cos β − sin β ⎞ ⎞ cos β − sin β ⎞
−1
cos β − sin β ⎞
T
⎛ ⎛ cos β − sin β ⎞ ⎞
⎜ ⎜ ⎟⎠ 0 ⎟ ⎜ ⎜ ⎟⎠ 0 ⎟ ⎜ ⎜ ⎟⎠ 0 ⎟ ⎜ ⎜ ⎟⎠ 0 ⎟
⎜ ⎝ sin β cos β ⎟ ⎜ ⎝ sin β cos β ⎟ ⎜ ⎝ sin β cos β ⎟ ⎜ ⎝ sin β cos β ⎟
⎜ ⎟ =⎜ ⎟ =⎜ ⎟ =⎜ ⎟
−1
⎛ ⎞ ⎟ ⎛ ⎞ ⎟
T
⎜ cos β − sin β ⎜ ⎛ cos β − sin β ⎞ ⎟ ⎜ ⎛ cos β − sin β ⎞ ⎟ ⎜ cos β − sin β
⎜ 0
⎜⎝ ⎟⎠ ⎟ ⎜ 0
⎜⎝ ⎟⎠ ⎟ ⎜ 0
⎜⎝ ⎟⎠ ⎟ ⎜ 0
⎜⎝ ⎟⎠ ⎟
⎝ sin β cos β
⎠ ⎜⎝ sin β cos β ⎟⎠ ⎜⎝ sin β cos β ⎟⎠ ⎝ sin β cos β

−1 T
Therefore, T (e) = T (e) , and the local-to-global transformation is orthogonal.
In summary, the order-4 local-to-global transformation equations for a truss
member are,

r ( ) = T ( )r ′ ( ) r ′ ( ) = T ( )T r ( )
e e e e e e

or
d = T ( )d′ d′ = T ( )T d
e e

Nodal Displ’s. Nodal


4 dof Two-Node Truss Element Actions

⎧ diX ⎫ 1 ⎧ riX ⎫
⎪ ⎪ ⎪ ⎪
⎪ diY ⎪ 2 ( e) ⎪ riY ⎪
d=⎨ ⎬ r =⎨ ⎬
⎪ d jX ⎪ 3 ⎪rjX ⎪
⎪d ⎪4 ⎪r ⎪
⎩ jY ⎭ ⎩ jY ⎭

Section 3-3. Two-node Truss Member Equations r ( ) = k ( )d . This two-node,


e e

planar truss element is summarized in the above Element Box. In Abaqus/CAE, it is


designated type T2D2 (or type T3D2 in three-dimensions).
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-11

Fundamentally, stiffness is defined as force divided by deflection in the direction


of the force. Therefore, the stiffness relationship for a truss member will depend on
which forces and deflections are used in the definition; different element stiffness
relationships will occur for each of the three coordinate systems presented in Section 3-2.

The member stiffness relationship in the global system, r ( ) = k ( )d , is most important.


e e

Its derivation may be based on either equilibrium or energy. Both derivations are
presented in the following.
3-3.1 Stiffness Relationship in Basic Coordinate System. The following basic
member stiffness relationship is introduced in elementary solid mechanics,
EA
F ( e) = Δ
L

Basic member stiffness terms

Therefore, in the basic coordinate system the truss member stiffness is simply,
F ( e) EAΔ EA
k ( e) =
Δ
=

=
L
(basic member stiffness )
3-3.2 Stiffness Relationships in Local Coordinate System.
From member equilibrium,

∑F x = 0 ⇒ rix′ = − rjx′ and ∑F y = 0 ⇒ riy′ = − rjy′ = 0

Therefore,
FBD of node j
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-12

(
rix′ = − rjx′ = −k ( ) d ′jx − dix′
e
)
(
= k ( ) dix′ − d ′jx
e
)
Also, from equilibrium of node j,
rjx′ = F ( e) . Therefore,

rjx′ = F ( ) =
e EA
L
Δ = k ( ) d ′jx − dix′
e
( )
(
= −k ( ) dix′ − d ′jx
e
)
The last two equilibrium equations may be
Nodal displacements and forces for the truss
expressed in the order-4 matrix system, member in the local coordinate system

⎧ rix′ ⎫ ⎧ d′ ⎫
⎪ ⎪ ⎛ 1 0 −1 0 ⎞ ⎪
ix

riy′ ⎪⎪
⎪⎪ ⎜ 0 0 0 ⎟ ⎪⎪ iy′ ⎪⎪
d
( e) 0 ( e) ( e)
⎨ ⎬= k ⎜ ⎟⎨ ⎬ or r ′ = k ′ d′
⎪ rjx′ ⎪ ⎜ −1 0 1 0 ⎟ ⎪ d ′jx ⎪
⎪ ⎜⎝ 0 0 0 ⎟⎠ ⎪ d ′
rjy′ ⎪ 0 ⎪
⎪⎩ ⎪⎭ ⎪⎩ jy ⎪⎭

Thus, k ′( ) denotes the truss member stiffness matrix referred to the local coordinate
e

system.
Alternatively, the total potential energy for the member is,
Π ( e) = U ( e) + V ( e)
Since strain is uniform in a truss
member, the member stain energy is,
L 2
( e) 1 ⎛ du ⎞
U = ∫ EA ⎜ ⎟ dx
0
2 ⎝ dx ⎠
⎛ Δ⎞
2
L
= EA ⎜ ⎟
2 ⎝ L⎠
1
= k ( e) Δ 2
2 Local system terms needed for
energy approach
1
(
U ( e) = k ( e) d ′jx − dix′ )
2

2
By definition of a truss structure idealization, transverse loads are not admitted, i.e., q(x)
= 0. Therefore, member potential energy is associated only with the two nodal axial
forces rix′ and rjx′ , i.e.,
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-13

(
V ( e) = − rix′ dix′ + rjx′ d ′jx )
The total potential energy of the member is therefore,

Π ( e) =
1 ( e)
( )
2
k d ′jx − dix′ − rix′ dix′ − rjx′ d ′jx
2
Invoking the principle of stationary potential energy,
∂Π ( e)
∂ ( dix′ )
( )
= −k ( e) d ′jx − dix′ − rix′ = 0 ⇒ rix′ = k ( e) dix′ − d ′jx ( )
∂Π ( e)
( )
∂ d ′jx
( )
= k ( e) d ′jx − dix′ − rjx′ = 0 ⇒ rjx′ = −k ( e) dix′ − d ′jx ( )
The matrix form of these two equations is,
⎧ rix′ ⎫ ( e) ⎛ 1 −1⎞ ⎧ dix′ ⎫
⎨ ⎬= k ⎜ ⎨ ⎬
⎩rjx′ ⎭ ⎝ −1 1 ⎟⎠ ⎩d ′jx ⎭

Observe how the energy method guarantees that the stiffness matrix is symmetric. The

element stiffness matrix derives from strain energy U ( e) , and the symmetry stems from

the quadratic form of U ( e) .


Inserting the placeholder quantities diy′ , d ′jy and riy′ = − rjy′ = 0 , while preserving

symmetry, results in the same order-4 matrix equation r ′( ) = k ′( )d′ that was obtained
e e

previously using equilibrium.


3-3.3 Stiffness Relationship in Global Coordinate System. The nodal force and
nodal displacement components in the global coordinate system when arrayed in matrix
form are respectively:
⎧ riX ⎫ ⎧ diX ⎫
⎪ ⎪ ⎪ ⎪
⎪⎪ riY ⎪⎪ ⎪⎪ diY ⎪⎪
r( ) = ⎨
e
⎬ and d = ⎨ ⎬
⎪ rjY ⎪ ⎪ d jY ⎪
⎪ rjY ⎪ ⎪ d jY ⎪
⎪⎩ ⎪⎭ ⎪⎩ ⎪⎭

The stiffness matrix that relates r ( e) and d is the member stiffness matrix referred to the

global XY coordinate system, and is denoted k( e) . It is obtained by transforming the local


stiffness matrix k ′ (e) to the global system as follows,
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-14

e e e
( e e e
)
r ( ) = T ( )r ′ ( ) = T ( )k ′ ( )d′ = T ( )k ′ ( ) T ( ) T d
e e

Therefore, the truss member stiffness relationship referred to the global coordinate
system is,

r ( ) = k ( )d where k ( ) = T( )k ′( )T( )T
e e e e e e

Alternatively, this matrix equation represents the four scalar, nodal equilibrium
equations, two at each node, in the XY system.
The explicit form of the stiffness matrix for the truss member when referred to the
global XY coordinate system is obtained by evaluating the orthogonal transformation

T( )k ′( )T( )T . Observe that the orthogonal transformation will preserve symmetry for the
e e e

member stiffness matrix. The result is,

⎛ cos 2 β sin β cos β − cos 2 β − sin β cos β ⎞


⎜ ⎟
( e) ( e) ⎜ sin2 β − sin β cos β − sin2 β ⎟
k =k ⎜ ⎟
⎜ cos 2 β sin β cos β ⎟
⎜⎝ symm. sin2 β ⎟⎠

Here, k ( e) is the basic member stiffness, i.e., EA/L for member (e)), and β is the angle
between the global X-axis and the member x-axis reckoned positive when counter-
clockwise. This is a generic result, good for any truss member in the XY plane.
Section 3-4. Example - Finite Element Method for a Cantilevered Truss
Structure. An aluminum cantilevered truss structure is isolated as shown below. It has
six unknown reactions, two of which are "secondary" moment reactions. It also has an
unknown internal bar force in each of its six members plus “secondary” moments in each
member arising from the relatively rigid gusset plates indicated at the joints. Knowledge
of the two secondary moment reactions as well as secondary member moments internal to
the bars is obviated by the assumption that truss structures may be idealized as pin-
connected trusses. Thus, the gusset plates and the secondary moments they engender are
disregarded in the basic analysis of trusses.6

6
A USA Today article (Nov. 13, 2008) on the I-35W bridge collapse referenced the report “NTSB: Design
Errors Factor in 2007 Bridge Collapse,” by Frederic J. Frommer. The NTSB investigation concluded that
gusset plates had only one-half the required thickness, and that this design flaw led to the collapse.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-15

Cantilever truss structure

The pin-connected truss idealization leaves only four unknown force reactions (r
= 4) and six unknown internal bar forces (m = 6) for a total of ten unknowns (m + r =
10). There are five joints (n = 5) with two force equilibrium equations each, so there are
ten equations of equilibrium (2n = 10) available that can be used for solution. Therefore,
the actual truss structure shown when idealized as a pin-connected truss is statically
determinate. An elementary hand solution for the bar forces and reactions could be
pursed using the method of joints and the method of sections. However, a finite element
hand solution is the present objective.
Since FEM makes no distinction between statically determinate and statically
indeterminate structures, using this simple statically determinate truss to illustrate FEM
solution methodology loses no generality. A detailed, organized, finite element hand
solution is also warranted to illustrate the methodology. That the structure is statically
determinate also facilitates illustration of the important task of hand-checking the finite
element solution.
The objective of the analysis is to determine the deformation of the truss, i.e., the
joint displacements, as well as the external reactions at the supports and the basic bar
forces internal to the structure.
3-4.1 Preprocessing Phase. The resulting structural idealization shown below is
based on typical simplifying assumptions for truss structures. Moments at supports and
member internal end moments due to gusset plates are assumed secondary and ignored,
as discussed above.
Further, loads are restricted to concentrated forces at joints; no distributed or
concentrated forces are permitted on a member proper. Weight of each member is
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-16

generally distributed to the joints or considered negligible. These additional assumptions


and the frictionless pin assumption render internal moments in all members zero. All
members are therefore two-force bars.

Structural idealization of the cantilevered truss

It follows that the prescribed displacement boundary conditions for the structural
model are those indicated by the two pin supports in the structural idealization. These
boundary conditions restrain components of displacement at joints 1 and 3 in the global
coordinate directions. Each pin support also gives rise to two rectangular components of
the reaction force as shown.
Geometry of the model is obtained from the structural idealization. It is prepared
for input using the global coordinates of the five joints and in a consistent set of units as
indicated in Table 3-1. In this case those units are inches. Specific coordinate values will
depend on the choice for a datum and are omitted. Automated graphical tools are
generally used to draw a truss structure including its joints on a computer’s display using
such data. In any case, the global coordinates of each joint in a consistent set of units are
developed and stored internally in the computer for use in solution processing.

Table 3-1. Required geometrical input data from the structural idealization
Joint X Y
(in) (in)
1 ... ...
! ! !
5 ... ...
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-17

Upon describing the section properties of each member, i.e., modulus of elasticity
E and cross-sectional area A, and upon describing the two given vertical loads, all in a
consistent set of units, the description of the structural idealization is considered
complete. Observe that this description is independent of any particular method of
solution. The focus is exclusively on an accurate description of the structural
idealization. That is, presumably the structural idealization could be the basis for
solution by methods other than finite element.
The assumptions made in going from the actual cantilevered truss structure to its
structural idealization constitute modeling error. Nonetheless, engineering experience
with truss structures and their design standards indicate that this modeling error is
acceptable in support of design.7
3-4.2 Solution Phase. The solution phase is comprised of meshing and element
selection, assigning code numbers, and assembly and solution of the finite element
equations KD = Q.
3-4.2.1 Meshing and element type. The idealized truss structure naturally
subdivides into six members and five nodes as shown below in the structural model.

Upon selecting a two-node truss member, the basic member stiffness k ( e) = ( EA L )


( e)
is

calculated towards defining member stiffness matrices k (e) for each member.

6 members
5 nodes
6 active DOF
4 inactive DOF

Structural model of the cantilevered truss

The positive sense for the local x-axis must be known for each member. As
indicated in the sketch, it establishes the i and j nodes for each member. This information

7
To mitigate risk of structural failure and collapse, structural design of trusses also includes due
consideration of alternative load paths using redundant bars and reinforced joints, as well as other
measures.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-18

is effectively prescribed at the time code numbers are generated for each member, i.e., the
member code numbers needed for assembly of the structure stiffness matrix K and load
matrix Q. The positive x-axis is taken into account when the angle β between the global
X- and local x-axes for each member is calculated. That occurs during calculation of
member stiffness matrices k (e) .
3-4.2.2 Code number assignment and assembly. By definition, each joint in a
planar truss structural model has as many as two rectangular components of displacement
or DOF. There are four inactive DOF at support joints 1 and 3 due to the prescribed
pinned boundary conditions. There are six active DOF denoted D1 through D6 for the

structural model. They are shown numbered and in their positive directions at each
unconstrained joint, i.e., joints 2, 4, and 5. This numbering is arbitrary (as is joint
numbering), but note that at each joint the X-DOF comes first and the Y-DOF second.
The objective of the solution phase is to calculate D1 through D6 . Recall that to
know such values is to have just enough information to know the complete deformation
state of a structural model.
Member code numbers are determined largely by inspection while referring to the
labeled DOF in the sketch of the structural model. The results are given in Table 3-2.
Each code number sequence is written consistent with an arbitrarily chosen direction for
the member’s positive x-axis shown on a structural model’s sketch. Or vice versa
depending on which is chosen first, each positive x-axis is consistent with an arbitrarily
chosen code number sequence.
The zero integers contained in the code numbers for the first three members,
0012, 1200 and 0034, respectively, have the effect of implementing the fixed boundary
conditions at the supports (i.e., D0 X = 0, D0Y = 0 ) that were prescribed in the
preprocessing phase. They define the sense of the positive x-axis for these members.
Writing instead 1200, 0012 and 3400, respectively, would have reversed the positive x-
axis direction for these three members.
It’s good practice where applicable to also include the angle β in a code number
table. This angle is used in the calculation of a truss member’s stiffness matrix, and
sequencing of the integers that comprise code numbers completes its definition. Recall
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-19

that β is positive when turned counter-clockwise from the X-axis. For example, because
the code number for member (4) is 1234 instead of 3412, the angle β in member (4)’s
stiffness matrix is positive 90° as shown in Table 3-2 instead of negative 90° . This is
because the code number 3412 would effectively reverse the positive x-axis for member
(4).
Table 3-2 Member code numbers for the structural model
Member Code # β
(e) i j deg.
(1) 00 12 0
(2) 12 00 143.1
(3) 00 34 0
(4) 12 34 90
(5) 12 56 0
(6) 34 56 -36.87

In any hand solution it is recommended that member stiffness matrices k ( ) be


e

written with code numbers placed adjacent to them as shown below to facilitate
subsequent assembly into the global stiffness matrix K .
Member stiffness matrices k( e) follow (see Section 3-3):

Member (1)
(1)
(1) ⎛ EA ⎞ 10 × 10 3 × 5.6
β = 0° k =⎜ = = 388.888 kip / in
⎝ L ⎟⎠ 144
0 0 1 2
0 1 0 −1 0
(1)
k = 0 0 0 0 0 × 388.888 kip / in
1 −1 0 1 0
2 0 0 0 0
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-20

Member (2)
( 2)
( 2) ⎛ EA ⎞ 10 × 10 3 × 1.25
β = 143.13° k =⎜ = = 69.4444 kip / in
⎝ L ⎟⎠ 180
1 2 0 0
1 0.64 −0.48 −0.64 0.48
k( ) =
2
2 −0.48 0.36 0.48 −0.36 × 69.4444 kip / in
0 −0.64 0.48 0.64 −0.48
0 0.48 −0.36 −0.48 0.36
Member (3)
(3)
(3) ⎛ EA ⎞ 10 × 10 3 × 0.4
β = 0° k =⎜ = = 27.7777 kip / in
⎝ L ⎟⎠ 144
0 0 3 4
0 1 0 −1 0
(3)
k = 0 0 0 0 0 × 27.7777 kip / in
3 −1 0 1 0
4 0 0 0 0
Member (4)
( 4)
⎛ EA ⎞ 10 × 10 3 × 1.2
β = 90° k ( 4) = ⎜ = = 111.111 kip / in
⎝ L ⎟⎠ 108

1 2 3 4
0 0 1 0 0
(4 )
k = 2 0 1 0 −1 × 111.111 kip / in
3 0 0 0 0
4 0 −1 0 1
Member (5)
(5)
(5) ⎛ EA ⎞ 10 × 10 3 × 1.6
β = 0° k =⎜ = = 111.111 kip / in
⎝ L ⎟⎠ 144
1 2 5 6
1 1 0 −1 0
(5)
k = 2 0 0 0 0 × 111.111 kip / in
5 −1 0 1 0
6 0 0 0 0
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-21

Member (6)
(6 )
(6 ) ⎛ EA ⎞ 10 × 10 3 × 0.5
β = −36.87° k =⎜ = = 27.7777 kip / in
⎝ L ⎟⎠ 180

3 4 5 6
30.64 −0.48 −0.64 0.48
(6 )
k = 4 −0.48 0.36 0.48 −0.36 × 27.777 kip / in
5 −0.64 0.48 0.64 −0.48
6 0.48 −0.36 −0.48 0.36

The next step in the solution phase is to assemble the member equilibrium
equations into the structure equilibrium equations KD = Q . The symmetric structure
stiffness matrix K is a six-by-six array of stiffness coefficients since six active DOF
exist. These global DOF numbers are placed along side both the rows and columns of K
to facilitate looping through and parsing the element stiffness matrices k( e) for their
individual contributions to K .
The results are the member stiffness contributions listed in their appropriate
positions of the K matrix shown below. For example, entry K 11 would contain
contributions from only those member stiffness matrices having the integer 1 in their
code numbers. In this way, all members except (3) and (6) participate in contributing to
K 11 . In fact only members (1), (2) and (5) contribute to K 11 , since member (4)'s
contribution is zero. These contributions are 388.888 , 44.4444 (i.e., 0.64 × 69.4444 ),
and 111.111 , respectively. Their algebraic sum is K 11 = 544.443 . Likewise, only
members (1), (2), (4) and (5) participate in forming K 12 , because only their code
numbers contain both integers 1 and 2. However, only member (2)’s contribution is
nonzero ( −0.48 × 69.4444 = −33.3333 ). The remaining entries are determined similarly,
while observing the symmetry of K .
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-22

1 2 3 4 5 6
388.888
1 44.4444 −33.3333 0 0 −111.111 0
111.111
25.0000
2 0 −111.111 0 0
111.111
27.7777
K= 3 −13.3333 −17.7777 13.3333
17.7777
111.111
4 symm. 13.3333 −10.0000
10.0000
111.111
5 −13.3333
17.7777
6 10.0000

The structure load matrix is defined by inspection and given as,


T T
Q = ⎢⎣Q1 Q2 Q3 Q4 Q5 Q6 ⎥⎦ = ⎢⎣0 −9 0 0 0 −6 ⎥⎦ kip

Recall that Q = Qconc + Qdist . In general, member code numbers are also used to

assemble distributed force matrices fq( e) into the structure load matrix Qdist (see Section 1-

2.2.3). For truss structures, however, loads are restricted to concentrated forces at the
joints and distributed forces on members are not admitted (i.e., q(x) = 0 ⇒ fq( e) = 0 ). It

follows for truss structures in general that Qdist = 0 and assembly of the matrix Q ≡ Q conc
is trivial. In the present hand solution illustration it is obtained largely by inspection with
due consideration for direction and sign of concentrated forces at the joints.
The matrix of global displacements for the structural model is defined as the
column matrix (alternatively, structural displacement matrix or DOF matrix),
T
D = ⎢⎣ D1 D2 D3 D4 D5 D6 ⎥⎦
Zero-displacement boundary conditions at supports of a structure are automatically
imposed with the proper choice of active DOF. During the identification and numbering
of active DOF the mere exclusion of the inactive DOF at joints, in this case joints 1 and
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-23

3, has the effect of imposing the zero-displacement boundary conditions. Thus, the
essential boundary conditions for the structure are satisfied explicitly in the solution.
The resulting finite element equilibrium equations for the structure KD = Q are,

⎛ 544.444 −33.3333 0 0 −111.111 0 ⎞ ⎧ D1 ⎫ ⎧ 0 ⎫


⎜ 136.111 0 −111.111 0 0 ⎟ ⎪ D ⎪ ⎪ −9 ⎪
⎜ ⎟ ⎪ 2⎪ ⎪ ⎪
⎜ 45.5554 −13.3333 −17.777 13.3333 ⎟ ⎪ D3 ⎪ ⎪ 0 ⎪
⎜ ⎨ ⎬= ⎨ ⎬
121.111 13.3333 −10.0000 ⎟ ⎪ D4 ⎪ ⎪ 0 ⎪
⎜ ⎟
⎜ symm. 128.889 −13.3333 ⎟ ⎪ D5 ⎪ ⎪ 0 ⎪
⎜ ⎟⎪ ⎪ ⎪ ⎪
⎝ 10.0000 ⎠ ⎩ D6 ⎭ ⎩−6 ⎭

3-4.2.3 Solution of the finite element equations. The expression D = K −1Q is a


symbolism for various implementations of the Gaussian elimination method used to solve
a linear systems of algebraic equations KD = Q. These numerical solution procedures
are highly optimized for very large systems in FEM software, and are categorized as
direct equation solvers. Again, the inverse global stiffness matrix K −1 is rarely if ever
calculated in FEM.
In this case, the solution D = K −1Q gives the following joint displacements of the
cantilevered truss structure. Positive and negative values correspond to positive and
negative DOF directions, respectively, which are shown at the joints in the previous
sketch of the structural model.
T
D = ⎢⎣-0.0719998 -0.696000 0.288000 -0.750000 -0.143999 -1.92600 ⎥⎦ inches
One rule of thumb in finite element analysis is to give the computed displacement
data D a sanity check prior to or at the beginning of the post-processing phase. Since
element forces and stresses follow from these data there is no point to proceeding if the
displacements are incorrect from cursory inspection.
In practice, the computer displays the shape of the deformed configuration of the
structure. That the shape is or is not consistent with applied loading can constitute a
sanity check. However, since magnitudes of the displacement are scaled to facilitate
display they too should also be rationalized. Here, the computed displacements do pass a
sanity check. For example, it is reasonable that the largest nodal displacement of the
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-24

cantilevered truss structure should be D6 in the negative direction, which it is, since it is
the tip deflection of a transversely loaded cantilever. The other nodal displacements D1
through D5 are consistent with this observation in both magnitude and direction.
It is well to appreciate at this point that the structural problem is essentially
resolved whether the structure is statically determinate or indeterminate. The nodal
displacements are the primary variables in the displacement-based finite element method,
and when they are in hand the complete deformation of the structure is described. In this
sense, the data processing that follows may be simply regarded as back-substitution for
the secondary variables, e.g., for the member basic forces and stresses.
3-4.3 Post-processing Phase. In this phase nodal displacement data D computed
in the solution phase are processed for selected secondary variable output. In this
instance, member basic forces F(e) are calculated to illustrate how the data are extracted
from the displacement matrix D and processed. This is followed by calculation of
reaction forces at the upper and lower pin supports for the subject truss. A general-
purpose computer method is discussed; a more tractable hand calculation method is
illustrated.
3-4.3.1 Basic Member Forces. The general equation used by the computer for
basic member forces F(e) is the equation for the member nodal forces r ′( e) in the local
coordinate system,

r ′ ( ) = T ( )T r ( )
e e e

= T ( ) T k ( )d
e e

where,
⎛ cos β sin β 0 0 ⎞
⎜ ⎟
( e)T ⎜ − sin β cos β 0 0 ⎟ T
T = and d = ⎢ diX diY d jX d jY ⎥
⎜ 0 0 cos β sin β ⎟ ⎢⎣ ⎥⎦
⎜ ⎟
⎜⎝ 0 0 − sin β cos β ⎟⎠

The angle β and the stiffness matrix k ( ) are recalled from memory (or
e

regenerated). In addition, entries in the member displacement matrix d , which are


referred to the global coordinate system, must be identified with entries in the global
displacement matrix D. Member code numbers will provide the needed reverse mapping
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-25

to do this. Finally, results from the above equation are used to calculate the basic
member force using,

F ( ) = rjx′
e

and member axial stress σ ( e) = F ( e) A if desired.

Calculation of the six basic member forces F ( e) follows:

Member (1)

0 0 1 2 T
β =0 o
d= T =⎢ 0 0 D1 D2 ⎥
⎢ d diY d jX d jY ⎥ ⎣ ⎦
⎣⎢ iX ⎦⎥

⎧ rix′ ⎫ ⎛
⎪ ⎪ 1 0 0 0 ⎞⎛ 1 0 −1 0 ⎞ ⎧ 0 ⎫
riy′ ⎪ ⎜ 0 ⎟⎜ ⎪ ⎪⎪
⎪ 1 0 0 0 0 0 0 ⎟⎪ 0
⎨ ⎬=⎜ ⎟⎜ ⎟⎨ ⎬ × 388.888
⎪ rjx′ ⎪ ⎜ 0 0 1 0 ⎟⎜ −1 0 1 0 ⎟ ⎪ -0.0719998 ⎪
⎪ ⎜ ⎟⎠ ⎜⎝ 0 0 0 ⎟⎠ ⎪⎩ -0.696000 ⎪⎭
rjy′ ⎪ ⎝ 0 0 0 1 0
⎩ ⎭
r ′(1) = T(1)T k(1) d

Therefore,
⎧ rix′ ⎫ ⎧
⎪ ⎪ 28.0000 ⎫
⎪ riy′ ⎪ ⎪⎪ 0
⎪⎪
⎨ ⎬= ⎨ ⎬
⎪ rjx′ ⎪ ⎪ −28.0000 ⎪
⎪ rjy′ ⎪ ⎪⎩ 0 ⎪⎭
⎩ ⎭

The basic member force is

F ( ) = rjx′ = −28.0000 or F ( ) = 28.0000 kip C


1 1
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-26

Member (2)

1 2 0 0 T
β = 143.130 o d= T = ⎢ D1 D2 0 0 ⎥
⎢ d diY d jX d jY ⎥ ⎣ ⎦
⎣⎢ iX ⎦⎥

⎧ rix′ ⎫ ⎛
⎪ ⎪ −0.8 0.6 0 0 ⎞⎛ 0.64 −0.48 −0.64 0.48 ⎞ ⎧ -0.0719998 ⎫
riy′ ⎪ ⎜ −0.6 −0.8 ⎟⎜ ⎪ ⎪
⎪ 0 0 −0.48 0.36 0.48 −0.36 ⎟ ⎪ -0.696000 ⎪
⎨ ⎬=⎜ ⎟⎜ ⎟⎨ ⎬ × 69.4444
⎪ rjx′ ⎪ ⎜ 0 0 −0.8 0.6 ⎟⎜ −0.64 0.48 0.64 −0.48 ⎟ ⎪ 0 ⎪
⎪ ⎜ −0.6 −0.8 ⎟⎠ ⎜⎝ 0.48 −0.36 −0.48 0.36 ⎟⎠ ⎪⎩
rjy′ ⎪ ⎝ 0 0 0 ⎪⎭
⎩ ⎭

Therefore,
⎧ rix′ ⎫ ⎧
⎪ ⎪ −25.0000 ⎫
⎪ riy′ ⎪ ⎪⎪ 0
⎪⎪
⎨ =
⎬ ⎨ ⎬
⎪ rjx′ ⎪ ⎪ 25.0000 ⎪
⎪ rjy′ ⎪ ⎪⎩ 0 ⎪⎭
⎩ ⎭

The basic member force is,

F ( ) = rjx′ = 25.0000 kip T


2

Member (3)

0 0 3 4 T
β = 0o d= T =⎢ 0 0 D3 D4 ⎥
⎢ d diY d jX d jY ⎥ ⎣ ⎦
⎣⎢ iX ⎦⎥
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-27

⎧ rix′ ⎫ ⎛
⎪ ⎪ 1 0 0 0 ⎞⎛ 1 0 −1 0 ⎞ ⎧ 0 ⎫
riy′ ⎪ ⎜ 0 ⎟⎜ ⎪ ⎪⎪
⎪ 1 0 0 0 0 0 0 ⎟⎪ 0
⎨ ⎬=⎜ ⎟⎜ ⎟⎨ ⎬ × 27.7777
⎪ rjx′ ⎪ ⎜ 0 0 1 0 ⎟⎜ −1 0 1 0 ⎟ ⎪ 0.288000 ⎪
⎪ ⎜ ⎟⎠ ⎜⎝ 0 0 0 ⎟⎠ ⎪⎩ -0.750000 ⎭⎪
rjy′ ⎪ ⎝ 0 0 0 1 0
⎩ ⎭

Therefore,
⎧ rix′ ⎫ ⎧
⎪ ⎪ −8.00000 ⎫
⎪ riy′ ⎪ ⎪⎪ 0
⎪⎪
⎨ ⎬= ⎨ ⎬
⎪ rjx′ ⎪ ⎪ 8.00000 ⎪
⎪ rjy′ ⎪ ⎪⎩ 0 ⎪⎭
⎩ ⎭

The basic member force is,

F ( ) = rjx′ = 8.00000 kip T


3

Member (4)

1 2 3 4 T
β = 90 o d= T = ⎢ D1 D2 D3 D4 ⎥
⎢ d diY d jX d jY ⎥ ⎣ ⎦
⎣⎢ iX ⎦⎥

⎧ rix′ ⎫ ⎛
⎪ ⎪ 0 1 0 0 ⎞⎛ 0 0 0 0 ⎞ ⎧ -0.0719998 ⎫
riy′ ⎪ ⎜ −1 ⎟⎜ ⎪ ⎪
⎪ 0 0 0 0 1 0 −1 ⎟ ⎪ -0.696000 ⎪
⎨ ⎬=⎜ ⎟⎜ ⎟⎨ ⎬ × 111.111
⎪ rjx′ ⎪ ⎜ 0 0 0 1 ⎟⎜ 0 0 0 0 ⎟ ⎪ 0.288000 ⎪
⎪ ⎜ 0 −1 ⎟⎠ ⎜⎝ 0 −1 0 1 ⎟⎠ ⎪⎩ -0.750000 ⎭⎪
rjy′ ⎪ ⎝ 0 0
⎩ ⎭

Therefore,
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-28

⎧ rix′ ⎫ ⎧
⎪ ⎪ 6.00000 ⎫
⎪ riy′ ⎪ ⎪⎪ 0
⎪⎪
⎨ =
⎬ ⎨ ⎬
⎪ rjx′ ⎪ ⎪ −6.00000 ⎪
⎪ rjy′ ⎪ ⎩⎪ 0 ⎪⎭
⎩ ⎭

The basic member force is,

F ( ) = rjx′ = −6.00000 or F ( ) = 6.00000 kip C


4 4

Member (5)

1 2 5 6 T
β = 0o d= T = ⎢ D1 D2 D5 D6 ⎥
⎢ d diY d jX d jY ⎥ ⎣ ⎦
⎣⎢ iX ⎦⎥

⎧ rix′ ⎫ ⎛
⎪ ⎪ 1 0 0 0 ⎞⎛ 1 0 −1 0 ⎞ ⎧ -0.019998 ⎫
riy′ ⎪ ⎜ 0 ⎟⎜ ⎪ ⎪
⎪ 1 0 0 0 0 0 0 ⎟ ⎪ -0.696000 ⎪
⎨ ⎬=⎜ ⎟⎜ ⎟⎨ ⎬ × 111.111
⎪ rjx′ ⎪ ⎜ 0 0 1 0 ⎟⎜ −1 0 1 0 ⎟ ⎪ -0.143999 ⎪
⎪ ⎜ ⎟⎠ ⎜⎝ 0 0 0 ⎟⎠ ⎪⎩ -1.92600 ⎪⎭
rjy′ ⎪ ⎝ 0 0 0 1 0
⎩ ⎭
Therefore,
⎧ rix′ ⎫ ⎧
⎪ ⎪ 7.99993 ⎫
⎪ riy′ ⎪ ⎪⎪ 0
⎪⎪
⎨ ⎬= ⎨ ⎬
⎪ rjx′ ⎪ ⎪ −7.99993 ⎪
⎪ rjy′ ⎪ ⎪⎩ 0 ⎪⎭
⎩ ⎭

The basic member force is,

F ( ) = rjx′ = −7.99993 or F ( ) = 7.99993 kip C


5 5
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-29

Member (6)

3 4 5 6 T
β = −36.870 o d= T = ⎢ D3 D4 D5 D6 ⎥
⎢ d diY d jX d jY ⎥ ⎣ ⎦
⎣⎢ iX ⎦⎥

⎧ rix′ ⎫ ⎛
⎪ ⎪ 0.8 −0.6 0 0 ⎞ ⎛ 0.64 −0.48 −0.64 0.48 ⎞ ⎧ 0.288000 ⎫
riy′ ⎪ ⎜ 0.6 0.8 ⎪ ⎪
⎪ 0 0 ⎟ ⎜ −0.48 0.36 0.48 −0.36 ⎟ ⎪ -0.750000 ⎪
⎨ ⎬=⎜ ⎟⎜ ⎟⎨ ⎬ × 27.7777
⎪ rjx′ ⎪ ⎜ 0 0 0.8 −0.6 ⎟ ⎜ −0.64 0.48 0.64 −0.48 ⎟ ⎪ -0.143999 ⎪
⎪ ⎜ 0.6 0.8 ⎟⎠ ⎜⎝ 0.48 −0.36 −0.48 0.36 ⎟⎠ ⎪⎩ -1.92600 ⎪⎭
rjy′ ⎪ ⎝ 0 0
⎩ ⎭

Therefore,
⎧ rix′ ⎫ ⎧−9.99995 ⎫
⎪ r′ ⎪ ⎪ ⎪
⎪ iy ⎪ ⎪ 0 ⎪
⎨ ⎬= ⎨ ⎬
⎪rjx′ ⎪ ⎪ 9.99995 ⎪
⎪⎩rjy′ ⎪⎭ ⎪⎩ 0 ⎪⎭

The basic member force is,

F ( ) = rjx′ = 9.99995 kip T


6

Our goal is to always emphasize how the computer does things so long as it can
be done with matrices that are amenable to hand calculation as in this case. However, it
should be noted in passing that an alternative hand calculation method sans matrices

(
using the formula F ( e) = k ( e) d jX − diX ) cos β would have been more expedient; it may
be deduced from Section 3-3.2. Element code numbers would be used to extract values
for d jX and diX from D in the manner illustrated above.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-30

3-4.3.2 Reactions. A general-purpose finite element procedure for calculating


reactions is based upon the unreduced nodal equilibrium equation for the structure,
R + = K + D + − Q conc+ . Here, R+ is the unreduced reaction force matrix and K + and Qconc+
are the unreduced structure stiffness and load matrices, respectively. In this case the
matrix equation is order-10 since six active DOF and four inactive DOF form the
unreduced structure displacement matrix D+. In a post-processing phase the right-hand
side of the equation is known, and reactions arrayed in R+ may be requested as output
data. In this case the six entries corresponding to the six active DOF would be null. In
FEM the number of null entries generally far and away exceeds the number of reactions.
The general-purpose procedure for reactions is therefore not suited to our purposes and is
omitted here in favor of a fundamental equilibrium analysis involving minimal hand
calculation.8
Structure reactions R1 through R4 , which act at joints 1 and 3, are calculated
from joint FBDs using known member basic forces as follows.

Reactions R1 and R2 from FBD of joint 1:

∑F X = 0 ⇒ R1 = 28.0000 kip

∑F Y = 0 ⇒ R2 = 0

Reactions R3 and R4 from FBD of joint 3:

∑F X = 0 ⇒ R3 = −28.0000 kip

∑F Y = 0 ⇒ R4 = 15.0000 kip

In a general-purpose FEM setting, member basic forces and reactions will accord in this
respect.

8
The general-purpose procedure for computing displacements and reactions is reserved for Chapter 6.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-31

Another rule of thumb in finite element analysis calls for an independent hand
check of overall structure equilibrium. This check uses the reaction data given by a finite
element computer program along with the intended load data. It is illustrated for the
present case by applying conditions of static equilibrium to the free body diagram of the
entire truss structure. Reactions R1 through R4 are given output data in the post-
processing phase, and the intended loads are 9 kip and 6 kip downward.
Invoking the three conditions of planar equilibrium for the FBD of the idealized
truss structure (Section 3-4.1):
(1) ∑ F = R + R = 28.0000 − 28.0000 ≈ 0 kip
X 1 3

(2) ∑ F = R + R − 9 − 6 = 15.0000 + 0 − 9 − 6 ≈ 0 kip


Y 2 4

(3) ∑ M = R × 108 + 9 × 144 + 6 × 288 = −28.0000 × 108 + 9 × 144 + 6 × 288 ≈ 0 kip − in


1 3

In FEM, one should expect that the resultant force acting on the structure is zero to within
machine precision as suggested here. If it is not, the prescribed loading is likely not what
was intended.
In addition to the reasonableness of the computed displacement data observed
earlier, computed reactions are shown to be in accord with overall equilibrium of the
truss structure. Therefore, it may be concluded preliminarily that the finite element
analysis of the truss structure has been conducted correctly. It is also advisable to spot-
check the output list of member basic forces. As demonstrated for this case in Example
3-1, the analyst accomplishes this by independently computing basic member forces
using the method of joints and the method of sections.

Section 3-5. Skew Boundary Conditions and Implementation. Occasionally


boundary conditions are prescribed in a joint coordinate system x ′y′ . Let the joint
coordinate system make an angle φ with the global coordinate system XY , and consider
the following simple truss structure with a skew boundary condition at joint 2. The skew
boundary condition in this case is specifically that the displacement component normal to
the inclined support equals zero and the displacement component tangent to the support is
unrestrained.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-32

Idealized truss structure with skew boundary condition

Two methods for implementing a skew boundary condition are discussed. The
first is an approximate method that illustrates the essence of skew boundary conditions
and is also an example of the penalty method for imposing displacement constraints in
FEM. The second is an exact method that is based upon coordinate transformation.
Bearing in mind that FEM is an approximate method with several built-in sources of error
(see Section 1-8), the approximate method could be viewed as adding yet another source
of modeling error. However, the error that is introduced may be reduced to any desirable
level, and so the approximate and exact methods are regarded as equally accurate in
practice. Both methods are commonly found in general-purpose FEM programs.
3-5.1 Approximate penalty method. This method entails modifying the idealized
structure as shown below. Specifically, inserting a relatively rigid two-node truss
member (or rigid spring element) between joint 2 and the support has the effect of
imposing the skew boundary condition. This added member is oriented normal to the
support's incline or in the y′ -direction and assigned an arbitrarily high basic member

stiffness k ( e) as indicated. It coerces the condition that the displacement component in


the y′ -direction remains essentially zero. At the same time, it provides for unconstrained
displacement (albeit small) along the incline in the x ′ -direction. The skew boundary
condition is thereby implemented in an approximate manner, and at the expense of
adding one member to the structural model of the truss.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-33

Modified idealized truss structure

The following example shows that the basic stiffness k ( e) of an added member
needs to be about 4 to 6 orders of magnitude greater than the stiffness of a characteristic
truss member to achieve uniform computational accuracy.

===============================================================
Example 3-5. The model shown of a three-
member, four-joint truss structure indicates a skew
boundary condition along a slope φ = 30° at one
joint. A horizontal load P also exists at this joint.
All other joints are completely restrained.
Use the approximate penalty method to
calculate the rectangular components of
displacement D1 and D2 in the global coordinate
directions at the skew boundary joint. Express the
solution in terms of the displacement parameter
PL/EA.
The code number table shown here for the model is Element Code # β
constructed to begin the solution. It includes code number (1) 0012 0°
0012 for the added member (4), which is completely
(2) 0012 -45°
restrained at the lower end. With this code number, its angle
(3) 0012 -90°
β must be 120°. It will be apparent in the assembly process (4) 0012 120°
to follow that its code number and its angle β could
equivalently have been written 1200 and -60°, respectively.
Stiffness matrices for all four members along with their code numbers are:
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-34

0 0 1 2 0 0 1 2
0 1 0 -1 0 0 12 −1 2 −1 2 12
EA EA
k (1)
= 0 0 0 0 0 × k (2)
= 0 −1 2 1 2 1 2 −1 2 ×
L 2L
1 -1 0 1 0 1 −1 2 1 2 1 2 −1 2
2 0 0 0 0 2 1 2 −1 2 −1 2 1 2

0 0 1 2
0 0 1 2
0 0 0 0 0 0 14 − 3 4 −1 4 3 4
EA EA
k (3) =0 0 1 0 −1 × k (4) = 0 − 3 4 34 3 4 −3 4 ×α
L L
1 0 0 0 0 1 −1 4 3 4 14 − 3 4
2 0 −1 0 1 2 3 4 −3 4 − 3 4 34

In the added member stiffness matrix k(4), α is a dimensionless penalty parameter.


Upon assembly the reduced structure equations KD = Q are,

EA ⎜ ( ( )
1+ 1 2 2 + α 4 ) − (1 (2 2 ) + 3α 4) ⎞⎟ ⎧⎪⎨ D 1
⎫⎪ ⎧
⎬= ⎨
P ⎫

(( ) ) (1 (2 2 ) + 1+ 3α 4) ⎟⎠ ⎪⎩ D
⎜ ⎟
L ⎜ − 1 2 2 + 3α 4 2 ⎪⎭ ⎩⎪ 0 ⎭⎪

Solutions D = K −1Q are tabulated as follows for increasing orders of magnitude


of the penalty parameter α :
α D1 ( PL EA) D2 ( PL EA) D2 D1

0 0.792893 0.207107 0.261204


100 0.763686 0.285559 0.373923

101 0.726846 0.384514 0.529017

102 0.717313 0.410121 0.571746

103 0.716206 0.413094 0.576781

104 0.716094 0.413396 0.577293

105 0.716083 0.413426 0.577345

106 0.716081 0.413429


0.57735 ≈ 1 3
! ! !
!

1 3
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-35

For the given φ = 30° slope the exact solution for D2 D1 is tan 30° or 1 3 . It
is only achieved in the limit of the penalty parameter α as indicated in the table. With
α = 10 6 and finite word length computation, the approximate solutions for D1 and D2 are
shown underlined. Their ratio is exact to at least five significant figures. This level of
accuracy is greater than required in engineering practice.
===============================================================

3-5.2 Exact coordinate transformation method. A fixed pin support completely


restrains the rectangular components of displacement at joint 1 in the model shown
below. Hence, the two corresponding homogeneous boundary conditions are imposed
automatically using the code number assembly procedure (Section 3-4.2.2). Assuming
that this has been done so that the two inactive DOF at joint 1 are eliminated, a partially
reduced order-6 system of finite element equations is the result and may be denoted
K*D* = Q* . This system is not invertible because it is not informed by the skew
boundary condition at joint 2. The nodal displacement matrix D* includes both D1 and
D2 as active DOF, so rigid body rotation about joint 1 must still be restrained. This is
manifested in a singular stiffness matrix K* and no unique solution of this partially
reduced system is possible.

T T
D* = ⎢⎣ D1 D2 D3 D4 D5 D6 ⎥⎦ Q* = ⎢⎣0 0 0 0 P 0 ⎥⎦
K*D* = Q*
Partially reduced structural model prior to imposition of skew boundary condition

Before the skew boundary condition may be imposed it is necessary to transform


the system K*D* = Q* so that the two finite element equations for joint 2 reference the
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-36

joint coordinate system x ′y ′ while the other four equations reference the global system
XY . The transformation of displacement components shown at joint 2 in the previous
sketch is,
⎧⎪ D1′ ⎫⎪ ⎡ cos φ sin φ ⎤ ⎧⎪ D1 ⎫⎪ ⎧⎪ D1 ⎫⎪
⎨ ⎬= ⎢ ⎥⎨ ⎬ = T2 ⎨ ⎬
⎪⎩ D2′ ⎪⎭ ⎢⎣ − sin φ cos φ ⎥⎦ ⎪⎩ D2 ⎪⎭ ⎪⎩ D2 ⎪⎭
where T2 is the indicated joint-2 transformation matrix. It is noted that it is an orthogonal
transformation matrix.
Let Λ denote the following partitioned order-6 orthogonal transformation matrix
wherein each sub-matrix is an order-2 square matrix (e.g., I is the order-2 identity
matrix),
⎛ T2 0 0⎞
⎜ ⎟
Λ=⎜ 0 I 0⎟
⎜⎝ 0 0 I ⎟⎠

Using this transformation matrix, the partially reduced system K*D* = Q* is transformed
so that the equations for joint-2 displacement components D1 and D2 are replaced by the
equations for components D1′ and D2′ . The transformation is (see also Section 3-3.3),

K*D* = Q*
D* = ΛT D′
K* ΛT D′ = Q*

( ΛK Λ ) D′ = ΛQ
* T *

Let this new order-6 system be denoted K′D′ = Q′ where the transformed partially
reduced stiffness matrix is K′ = ΛK* ΛT , and the transformed partially reduced load
matrix is Q′ = ΛQ* . This system is also not invertible since the skew boundary condition
preventing rigid body rotation about joint 1 is still to be imposed.
The skew boundary condition is now imposed by letting D2′ = 0 and eliminating
the second equation of the system K′D′ = Q′ . This results in the fully reduced order-5
system KD = Q shown below, in which the DOF have been re-numbered 1 through 5 for
convenience. The structure stiffness matrix K is obtained by striking the 2nd row and 2nd
column of K′ , and the structure load matrix Q is obtained by striking the 2nd row of Q′ .
Since D2′ = 0 rigid rotation about joint 2 is now restrained. Therefore, K is non-singular
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-37

and the fully reduced system KD = Q is invertible for the structure’s displacement
components D.

T T
D = ⎢⎣ D1′ D2 D3 D4 D5 ⎥⎦ Q = ⎢⎣0 0 0 P 0 ⎥⎦
KD = Q
Fully reduced structural model with imposed skew boundary condition

It should be noted that to compute basic member forces in a post-processing


phase, computed skew displacements such as D1′ must first be resolved into rectangular
components in the global XY coordinate system.
The exact coordinate transformation method for imposing skew boundary
conditions, similar to what has been described here, is intended for implementation in
general-purpose finite element programs. Sophisticated algorithms constructed around
elementary row/column operations from linear algebra are employed. They address
multiple sets of rows and columns, not necessarily adjacent, in the structure stiffness and
load matrices of global systems of finite element equations. Procedures are repeated on
the fly for each set, i.e., for each skew boundary condition.
3-5.3 Closing Remarks on Skew Boundary Methods. Both the approximate
penalty method and the exact coordinate transformation method are generally applicable
to structures other than truss structures in FEM. Other methods have been proposed and
implemented for imposing skew boundary conditions, but these two methods were
preferred early on.9 They are also useful for implementing displacement constraint
equations in FEM. In nonlinear settings the approximate penalty method offers a
measure of convergence control through adjustment of the penalty parameter.

9
Gallagher, R. H. (1975). Finite Element Analysis Fundamentals, Prentice-Hall, Inc., Englewood Cliffs,
New Jersey, 420 pages, 1975.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-38

Section 3-6. /CAE Study - Saint Venant's Principle for Truss Members.
Saint Venant’s principle applied to the truss member shown states that axial stress
can be assumed uniformly distributed
according to elementary theory across
width w at a distance d from the point of
application of load P, as long as d is
equal to or greater than w. Otherwise,
within the range defined by the
characteristic dimension d axial stress is statically indeterminate and non-
uniformly distributed as indicated in the sketch.
Objective: The objective of this example is to study the substance of St. Venant’s
principle for a truss member subjected to a system of concentrated loads resembling the
effect of a pin and a gusset plate after their removal from the end of the member. To do
this, a plane stress finite element analysis of the member is
conducted using Abaqus/CAE.10
Subject Model: Relevant properties and loading for a
notional, vertical truss member are given below. The base
of the member is assumed restrained against vertical
displacement. Distance d is measured from the bottom of
the hole, which is considered a geometric discontinuity.
The pin force P1 is the main load, but other "gusset" loads
P2, P3, P4 may also participate. Here, the resultant axial load
P on the member is taken to be,
P = P1 + 2P3 sin 45 ! + P4
Dimensions (m) Loads (N) Material Properties
Length, L=0.91 P1=79360 E=180 GPa
Width, w=0.1524 P2=3120 ν=0.3
Thickness, h=0.003175 P3=8820
Radius, r=0.0381 P4=4410

According to elementary theory, while normal stress σ y in a bar is in general


statically indeterminate, the average normal stress σ avg over the cross-sectional area A at
any point along the length of the bar is easily computed from statics as,

10
Plane stress finite element analysis is studied in Chapter 9.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-39

F
σ avg =
A
where F is the internal bar force and equals the axial load P.
However, St. Venant's principle states that for a distance d (assumed
here to be measured from the bottom of the hole) that is equal to or
greater than the width of the bar w, the normal stress is statically
determinate and given by elementary theory as,
P
σy = (d ≥ w)
A
Otherwise, for d < w (and also for regions about and above the hole),
the normal stress σ y in the bar remains statically indeterminate.
This normal stress is computed for the subject
truss member using a finite element analysis of
its structural idealization shown above.
Results: The figure to the right
indicates the model's mesh, shows the contours
of predicted normal stress σ y , and indicates a
typical station below the hole along which
normal stress is sampled for plotting purposes.
Stresses in the d-region (i.e., d<w) and also above and around the hole given by
the finite element analysis are statically indeterminate, i.e., not available from elementary
bar theory. However, beyond the d-region, St. Venant's Principle states that the normal
stresses σ y are indeed given by the above elementary theory formula.
This is corroborated in the set of curves plotted below in which the finite element
data markers are taken from points along six horizontal stations below the hole, which are
defined in the legend. The ordinate is the normal stress σ y normalized on the average
normal stress σ avg = F A . The abscissa is the normalized distance across the width of
the bar x/w. It is clear form these results that elementary bar theory does not apply for
ratios d/w less than one. In this region the normal stress is given correctly only by finite
element data. However, where d/w ≥ 1 the uniform distribution of normal stress given by
elementary bar theory is accurate.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-40

Conclusion: Finite element analysis of a truss member subjected to axial load


serves to corroborate St. Venant's principle. Elementary theory is valid only for regions
that are located beyond a characteristic distance equal to the width of the bar from
applied loads and/or geometric discontinuities. This applies to the two-node truss
member presented in this chapter (which is Abaqus element type T2D2), since its
formulation is also based on elementary theory. The axial stress in a truss member
obtained from finite element analyses of truss structures, which is given by σ ( e) = F ( e) A ,
is valid for the member proper, but invalid in the d-region at either end of the member.
ANALYSIS OF PLANAR TRUSS STRUCTURES 3-41

/CAE Modeling Tips:


The following two /CAE
features were employed
in this study:
(1) Distributing a
concentrated axial load
using the Interaction
Module. A concentrated
load applied at the center
of a bolt hole can
simulate a truss pin force,
in this case the vertical tensile load P1 at the end of the
truss member. (The load P1 , not shown, is created in the
Load module.) It is distributed uniformly to the upper half
or bearing surface of the hole. First, a reference point (RP-2 as shown) is created at the
hole’s center. Second, by specifying a kinematic coupling constraint the bearing surface
is defined and mechanically coupled to the reference point in the Edit Constraint dialogue
shown. All nodes on this surface are constrained to displace in unison with the reference
point upon checking the DOF U1 and U2 as indicated. In effect a rigid cylindrical pin
bearing on the upper surface of the hole is simulated.
(2) Displaying stress contours in the Visualization Module. Stress contours can be
rendered and displayed in various ways. Here, it is shown in a sequence of three steps
using the Contour Plot Option dialogue how one may clearly display the contours.

(a) Choose a banded contour type (b) Show the contour edges

(c) Remove the mesh (choose No edges)


ANALYSIS OF PLANAR TRUSS STRUCTURES P3-1

PROBLEM SET 3
Problem 3-1
A two-bar truss structure is subjected to a vertical load P at the crown as shown. The
axial spring constant of each bar is k . Determine the vertical deflection δ of the crown
in terms of the load P by:
(a) Applying the conditions of equilibrium to a free body diagram of the crown joint,

(b) Applying the principle of stationary total potential energy to the structural system.

Problem 3-2
(a) What is a two-node truss element? Explain in words. How does it differ from the
two-node spring element discussed in Ch. 1?

(b) Identify and label the following on a sketch of the truss structure shown: its
displacement DOF (active and inactive) and its elements indicating their local positive x-
coordinate direction.

(c) Assemble the global stiffness and load matrices K and Q, respectively, using the
direct stiffness method (include a code number table). Assume all bars have the same
Young’s modulus E and cross-sectional area A.

(d) Solve for nodal displacement matrix D = K-1Q (in terms of the system’s parameters).
ANALYSIS OF PLANAR TRUSS STRUCTURES P3-2

Problem 3-3
For the truss structure finite element model shown:

(a) Identify and label the active DOF on a sketch of the


model.

(b) Find the stiffness matrix K and load matrix Q.

(c) Solve the finite element equations KD = Q for the


vertical deflection beneath the load P.

(d) Using the results from (c) calculate the potential


energy associated with the load P, i.e., Vp, and the work done by the load P, i.e., Wp.

Problem 3-4
(a) Sketch the finite element model of the
truss structure shown and on it label the
active DOF.

(b) Form the structure equations KD = Q for


the model. The following data pertain:

(e)
k ( e)
(1) 2k
(2) k
(3) k

(c) Solve for the structure displacements D in terms of the load P and stiffness k.

(d) Find the basic member forces F ( e) using the post-processing method of Section 3-4.3.

(e) Find the reaction forces using the results from part (d).

(f) Show reactions from (e) on a sketch of the structure. Check overall equilibrium of the
structure. Check the member forces computed in part (d) using the method of joints.

Problem 3-5
Each member of the truss structure shown has the same
stiffness k, and is 45° from the horizontal. Use the
finite element method as follows (hint: exploit
symmetry, and perhaps also superposition):
ANALYSIS OF PLANAR TRUSS STRUCTURES P3-3

(a) Identify and label the active DOF on a sketch of a finite element model of the
structure.
(b) Assemble the structure stiffness matrix K and load matrix Q.

(c) Solve the structure equations KD = Q and determine the horizontal and vertical
displacements of the load point.

(d) Determine the total strain energy U stored in the structure, and the total work W done
on the structure by the loads.

Problem 3-6
Exploit the symmetry of the truss structure described in Tutorial T3, and use the finite
element method as illustrated in Section 3-4 to:

(a) Solve for the displacement matrix D.

(b) Find the basic member forces F ( e) .

(c) Find the reaction forces at the supports.

(d) Verify the reactions from part (c) using overall equilibrium of the structure, and then
verify the member forces from part (b).

Problem 3-7
Solve for the following quantities using the exact coordinate transformation method
(Section 3-5.2) to address the skew boundary condition in the truss model shown in
Example 3-5 (Avoid any temptation to skew the global XY coordinate system 30° !):
(a) Horizontal deflection of load P in terms of the displacement parameter PL EA .

(b) Member forces F ( e) in terms of the loads.

(c) Reaction force R in terms of the load P.

Problem 3-8
Modify the truss structure shown in Problem 3-2 by rotating counterclockwise 30° both
the load P1 and its roller support thereby creating a skew boundary condition. Solve for
the following quantities using the exact coordinate transformation method (Section 3-5.2)
to address the skew boundary condition (Hint: Consider the nature of your solution when
P1 = P2):
(a) Displacements of P1 and P2 in terms of parameters P1 L EA and P2 L EA .
ANALYSIS OF PLANAR TRUSS STRUCTURES P3-4

(b) Member forces F ( e) in terms of loads P1 and P2.

(c) Reaction forces in terms of loads P1 and P2.

(d) Check overall equilibrium.

Problem 3-9
Redesign the subject steel truss structure of Tutorial T3 to reduce the maximum
deflection under the same load system without increasing the weight of the structure.
Describe the new design and document it with corroborating results from a re-analysis
using Abaqus/CAE. Include a printout of the structural idealization, the finite element
model with nodes and elements labeled, and the displaced shape. What is the new
deflection?

Problem 3-10
(a) For the subject cantilever truss of Section 3-4, use Abaqus/CAE to determine the
deflection of each joint, the reaction forces at the root, and the axial force and stress in
each member.

(b) Verify the results by first checking the reactions using overall equilibrium, and then
the forces in the members using the method of joints and/or the method of sections.

Problem 3-11
Solve the truss structure shown in Problem 3-7 using Abaqus/CAE. Provide print outs of
the finite element model with element and node numbers labeled, the deformed shape of
the model, and the report file from the analysis including nodal displacements, reactions,
and element stresses. Let E = 200GPa , A = 322 mm2 , L = 3658 mm , and P = 40 kN .
(Hint: To impose a skew boundary condition at a joint (e.g., a joint located at X=3658,
Y=0, Z=0), create a new XYZ coordinate system with origin at the joint and rotate it by
entering coordinates of a point through which the new X-axis runs (e.g., X=3658.866,
Y=0.5, Z=0, for a 30o-rotation). Then, prescribe the boundary condition in the usual
manner, but referred to the newly created joint coordinate system (e.g., prescribe U2=0,
for a roller support along the incline).

Problem 3-12
(a) A Howe roof truss subjected to an unsymmetrical snow load on the lee side is shown
below. All members have a cross-sectional area A = 5.0 in 2 and a modulus of elasticity
E = 29 × 10 3 ksi . Use Abaqus/CAE to calculate the deflections at each joint, the reaction
forces at the supports, and the stress in each member. Provide a plot of the deformed
shape, and printed lists of the reactions, displacements and member stresses.
ANALYSIS OF PLANAR TRUSS STRUCTURES P3-5

(b) Verify the results by checking the reactions using overall equilibrium, and then spot-
checking the forces in a few members using the method of joints and/or the method of
sections.

(c) Use Abaqus/CAE to calculate the deflections at each joint, the reaction forces at the
supports, and the stress in each member, for a symmetrical snow load (with a 4 kip load
applied at the crown). Provide a plot of the deformed shape, and printed lists of the
reactions, displacements and member stresses.

(d) If the stresses in part (a) are known to be accurate, demonstrate that these data can be
used along with the principle of superposition to spot check the stresses from part (c).

Problem 3-13
Consider a dead load stress analysis of the Warren bridge truss shown below. The dead
loads are 7.5 kips at each joint of the top chord, and 31.5 kips at each joint of the bottom
chord. The top chord has a cross-sectional area AT = 40.0 in 2 , the bottom chord has a
cross-sectional area AB = 20.0 in 2 , and the remaining members have a cross-sectional
area A = 30.0 in 2 . All members have modulus of elasticity, E = 29 × 10 3 ksi .

(a) Use Abaqus/CAE to calculate the joint deflections, the member stresses and the
reaction forces at the supports (recommended not exploiting symmetry). Provide a plot
of the model’s mesh labeled with element numbers, the deformed shape, and a printed list
of the reactions and member stresses.

(b) Verify the results by checking the reactions using overall equilibrium, and then spot-
checking the stresses in a few members using the method of joints and/or the method of
sections.
ANALYSIS OF PLANAR TRUSS STRUCTURES P3-6

Problem 3-14
Although ABAQUS has the capability of explicitly modeling both geometric and
constitutive nonlinearities, in this course we will only make use of the linear elastic
solution procedures. However, in practice there are a number of ways that an analyst can
approximate nonlinear behavior with a purely linear code. This problem illustrates one
such approach applied to buckling of truss bars under compression. As a side note, the
approach you will take here mirrors that of some analysts in the aerospace industry where
structures are sometimes designed to perform even after some members may buckle.
Consider a modification to the truss structure from tutorial T3, in which additional
diagonal members are added as shown in blue in part (a) of the figure. Note that the new
diagonal members cross over, but are not connected to the pre-existing members, so the
intersection points should not be considered as joints. (Of course to actually build such a
truss, the diagonal bars would have to be slightly offset in 3-D to avoid physical overlap,
but this is a subtle feature we can neglect in our analysis.)

(a) Pre-buckled model of the truss. (b) Post-buckled model of the truss.

From examination of the figure it seems clear that the new diagonal bars will be under a
compressive axial load given the direction of the applied forces F. As you know from
your previous course(s) on strength of materials, a simply supported bar under axial
compression will buckle when the axial force reaches a certain critical value. The goal of
this problem is to determine the response of the modified truss before and after buckling.

(a) Determine the critical value of the applied force F shown in the figure that will cause
the two new bars to buckle. To do this, first find the critical axial force for each of the
bars (call this Pcrit ) given the modulus and dimensions of the bars. Next, find the
relationship between the applied load F and the axial forces P in the bars by modifying
your ABAQUS/CAE analysis from the tutorial by adding one additional diagonal bar to
ANALYSIS OF PLANAR TRUSS STRUCTURES P3-7

your model of half of the symmetric structure. From this determine the critical force Fcr
that will cause the new bars to buckle.

(b) Determine the response of the truss if it is loaded past the critical point at which the
new bars buckle. For this purpose, let’s consider two similar models to describe the
nonlinear post-buckling response. First, let’s assume that immediately after the buckling
load is reached the buckled bars might also retain some load-carrying capacity (which
seems reasonable since it was stable before we added the new bars). Specifically, let us
assume that after the two bars buckle the load carried by each is fixed at the value P = Pcr
computed in Part 1, even as the total load F is increased further past Fcr . Determine the
post-buckled response of this model by replacing the new bars by the appropriate external
load P = Pcr applied to the joints, as shown in part (b) of the figure. Based on the results
of this analysis, sketch plots of F vs. the vertical deflection of the top and bottom joints (a
separate plot for each joint). In particular, indicate what the slopes of the force curves are
before and after buckling.

(c) On second thought, we might decide that it is a bit too risky to assume that the bars
won’t fail catastrophically when buckling. Accordingly, a more conservative model we
could consider is to assume that once the bars buckle, they simply vanish from the model,
and the structure subsequently responds as in Tutorial T3, i.e., as shown in part (b) of the
figure with P = 0 . Using the analysis results from that tutorial and from Part 1 of this
problem, sketch plots of F vs. the vertical deflection of the top and bottom joints (a
separate plot for each joint). In particular, indicate what the slopes of the force curves are
before and after buckling, along with the value of any discontinuous drops in force that
might occur as the buckling load is reached.

You might also like