Sei sulla pagina 1di 32

Economy of Steel

Framed Buildings
Through
Identification of
Structural Behavior
Flnley A. Charney
Author
Dr. Finley A. Charney earned both
his bachelor of science degree in
civil engineering and master of
science in architectural engineer-
ing from the University of Texas in
Austin, and his Ph.D. in engineer-
ing at the University of California at
Berkeley. His professional ex-
perience includes three years as a
design engineer at Walter P. Moore
and Associates in Houston, Texas,
four years as technical analyst at
Martin & Martin (formerly KKBNA)
in Denver, Colorado and four years
at his current position as associate
principal at J.R. Harris & Company
in Denver. He also serves as presi-
dent of Advanced Structural Con-
cepts, a software development and
advanced analysis subsidiary of
J.R. Harris & Company. Through
teaching and supervision of
graduate research, Dr. Charney
has maintained close ties with the
University of Colorado and the
Colorado School of Mines. He is a
professional engineer in the states
of Texas and Colorado.
Dr. Charney is an active mem-
ber of several technical societies
including the American Institute of
Steel Construction (AISC), the
American Concrete Institute (ACI),
the American Society of Civil En-
gineers (ASCE), and the Council
on Tall Buildings and Urban Habitat
(TBC). His current committee af-
filiations include the ASCE Com-
mittee on Tall Buildings, ASCE
Committee on Methods of
Analysis, ACI Committee 435
(Deflection), ACI Committee 442
(Lateral Loads), and TBC Commit-
tee 21-C (Elastic Analysis of Tall
Concrete Buildings).
Summary
Many engineers realize that
economy in steel frame building
construction may be achieved
through the identification of struc-
tural behavior. In this sense,
several computer programs are
available that identity those mem-
bers which are either too strong or
too weak, and by simple stress
ratios provide guidance as to which
member properly should be
modified to most economically
meet the governing strength re-
quirements.
For buildings that are drift or
perception of motion controlled,
traditional practice has left it to the
engineer's experience and judge-
ment to locate sources of excess
stiffness or flexibility. With the ad-
vent of newly available computer
programs however, the engineer's
abilities may be augmented by
precise information on deforma-
tional aspects of behavior. This in-
formation includes identification of
those beams, columns, braces,
and beam-column joints which are
either too flexible or too stiff,
together with detailed data on the
sources of excess stiffness or
flexibility within the members iden-
tified.
In this paper, one of the most
powerful methods of structural
deformation identification is
presented by theory and example.
This method, which is based on the
principle of virtual work, has been
used successfully by design firms
nationwide to increase the efficien-
cy and economy of steel building
structures.
12-1
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
ECONOMY OF STEEL FRAMED BUILDINGS
THROUGH IDENTIFICATION OF STRUCTURAL BEHAVIOR
INTRODUCTION
In the design of steel building structures, the engineer's primary goal is to select a structural scheme that works
within architectural and mechanical constraints, and proportion the individual members of the structure so that all
relevant serviceability and strength requirements are satisfied. The realization of this goal is complicated by demands
on the engineer's time, and by the basic requirement that the structure be economical.
When strength controls the design of the structure, the proportioning of the various members is relatively
straightforward, and is often facilitated by one of several available computer aided design programs. These programs
identify weak members, and provide suggestions for strengthening.
When stiffness controls the design, reproportioning of the members is not as easy because the available structural
analysis programs provide only a global index of stiffness. After the results of the analysis are reviewed, the
engineer knows that the structure is too flexible or too stiff, but has no information regarding the sources of excess
flexibility or stiffness. This is equivalent to the design program stating that the structure has insufficient strength,
and then refusing to identify the location and the nature of the overstress.
Given this predicament, the design engineer has little choice but to use intuition and experience as a basis for resizing
the members of the structure, re-analyze, and assess the results. This "traditional" design process, illustrated in Figure
1, will eventually produce a structure with sufficient stiffness. Unfortunately, the successful design will be
accomplished at the expense of the engineer's time, and without proper regard to the economy of the structure.
Over the past ten years, tall building designers have been using a technique that solves many of the problems
associated with proportioning structures for stiffness. This method, originally described by Velivasakis and DeScenza
(20), uses the principle of virtual work to identify sources of excess flexibility and stiffness in the structure. More
recent illustrations on the use of the technique have been presented by Baker (1,2), Brazil and Scarangello (4),
Charney (7-11), Forrest-Brown and Samali (13), Henige (16), lyengar and Sinn (17), McNamara (18), Thornton,
et al. (19), and Wada (21,22).
Using the information provided by the virtual work analysis, the engineer can quickly identify which members are
too flexible or too stiff, determine the source of excess flexibility or stiffness within the member, and then resize
the members in the most effective way possible. The impact of the virtual work method on the design process is
shown in Figure 2, where it may be seen that three new steps have been added to those shown in Figure 1. Each
of these new steps is described in subsequent sections of this paper. Before continuing, however, it should be noted
that the application of the virtual work method is not limited to tall buildings. It is equally applicable to low- and
mid-rise buildings, bridges, gantries, microwave antenna towers, or any other stiffness or vibration sensitive structure.
Interestingly, Balling (3) has shown that the virtual work methods already in use by tall building designers are deeply
embedded in the formal optimum design methodologies described by Grierson, et al. (5,14,15). Without being fully
aware of the more formal techniques, however, many users of the virtual work method of behavior identification have
also referred to their process as optimization.
12-3
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Figure 1. Flow Chart for Traditional Stiffness Controlled Design
Figure 2. Revised Flow Chart Incorporating Virtual Work and Behavior Identification
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
While computer programs based on formal optimization techniques have merit, and as stand-alone utilities are more
powerful than the virtual work method used alone, they have not been readily accepted by the design profession. The
lack of acceptance is primarily due to the fact that structural engineers like to be in control of the design process,
and as such, are reluctant to pass complete responsibility to a "black box." Another reason formal optimization has
not been fully utilized lies in the manner the method has been presented. Papers on the technique are usually
presented in terms of advanced mathematics, with little emphasis on what the various equations mean with respect
to structural behavior.
The virtual work method as used by building designers provides an elegant solution to this quandary, because the
mathematical basis of the technique is already familiar to most designers, and may be presented in terms of physical
behavior. Furthermore, the information required for optimization is made directly available to the designer, who may
then take full control of the reproportioning and redesign. The process of using virtual work to resize the members
of overly flexible building structures is referred to as "human-loop optimization".
In practice, most of the engineers using virtual work optimization have developed postprocessors that are closely
associated with a particular large-scale structural analysis program. These postprocessors read the results from "real"
and "virtual" load cases, and then print or plot results useful to the engineer. Some programs go beyond this, having
the additional ability to automatically resize members on the basis of a certain target stiffness. One such program
is DISPAR (7,8) which is designed to work with the commercially available programs ETABS and SAP90
1
.
OBJECTIVES and SCOPE
The main objective of this paper is to present, through theory and example, the virtual work method as used in the
resizing of the members of stiffness controlled building structures. While some of the material presented is new,
much of it has been previously published in other forms by the writer and by other engineers working in the area.
In the first part of the paper, the virtual work method is applied to truss structures, whose only source of deformation
is axial. The technique is then applied to frames that have axial, flexural, shear, and torsional deformation, and is
finally expanded to incorporate deformations occurring in the panel zones of beam-column joints. The paper ends
with recommendations for future research in virtual work optimization of steel structures.
THE VIRTUAL WORK METHOD AS APPLIED TO TRUSSES
Consider the simple braced truss of Figure 3. This structure, which is modelled as an assembly of pinned-ended
bars, is subjected to a lateral load pattern R. The length and cross sectional area of each member are known. The
objective of the analysis is to determine how much each member of the truss contributes to the displacement at the
top. To compute the lateral deflection at the top of the truss, the principle of virtual forces is invoked:
W
B
=W
1
(1)
where W
E
is the external work produced by a virtual force Q moving through the unknown real displacement 8, and
W, is the sum of the internal work done by each member whose real stresses move through virtual strains acting
through the volume of the member.
1
ETABS and SAP90 are registered trademarks of Computers and Structures, Berkeley, California.
12-5
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
The internal work for each member i is :
where is the axial force in the member due to the loads R, is the axial force in the member due to the virtual
force Q, is the member area, E is the modulus of elasticity, and v indicates that the integral is taken over the
volume of the member. For a member of length the integral reduces to:
(2)
(3)
Figure 3. Planar Truss Used for Illustration of the Virtual Work Method
12-6
A) REAL STRUCTURE AND
APPLIED LOADS
B) VIRTUAL FORCE
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
The internal work for all n members of the structure is:
On the basis that internal work equals external work
Although the magnitude of the virtual force is usually taken as unity, Q is retained in equation (5) for consistency
of units.
In equation (3), the quantity is called the member's displacement participation factor (DPF) because it represents
how much the member contributes to the real displacement in the direction of the virtual force. In the case of the
truss, the entire member DPF is due to axial deformation. To distinguish axial deformation sources from other
sources, the value represented by in equation (3) will later be referred to as indicating that it represents
the axial component of the displacement participation factor of member i.
As a numerical example, consider again the structure of Figure 3. It is desired to determine the contribution of each
member of the truss to the lateral deflection occurring at the top. Here, the "real loads" consist of three 10 kip lateral
forces, and the "virtual load" is a 1.0 kip force applied horizontally at the top. Note that each member has a cross-
sectional area of one square inch, and that the total volume of the structure is 2080 cubic inches.
The results of the virtual work analysis are shown in Table 1. Each member's DPF is given in column seven of the
table, and the sum of these values (11.39 inches) is the real displacement at the top of the truss and in the direction
of the virtual force. Deformations occurring in member-5 contribute the most to the total displacement, with a DPF
of 3.24 inches, and member-1 contributes the least with a DPF of only 0.14 inches.
While member-5 contributes more to the total displacement than any other member, it is not the most critical
member if the aim is to reduce deflection with a minimum volume of added material. The most critical such element
is member-3, which contributes 3.04 inches to the total displacement. The fact that member-3 is critical is evident
from the value shown in column eight of Table 1, which represents how much that member's DPF changes if the
member volume is increased by 10.0 cubic inches. (Note that the values in column eight have been multiplied by
1000 for formatting purposes only).
12-7
(5)
(4)
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
1
Member
1
2
3
4
5
6
7
8
9
10
TOTALS
2
Length
L(in)
240
240
120
240
240
200
200
200
200
200
-
3
Area
A (in
2
)
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
-
4
Volume
V(in
3
)
240.00
240.00
120.00
240.00
240.00
200.00
200.00
200.00
200.00
200.00
2080.00
5
Real
Force
P (kips)
7.50
30.00
67.50
-15.00
-45.00
-12.50
12.50
-25.00
25.00
-37.50
-
6
Virtual
Force
P (kips)
0.75
2.25
3.75
-1.50
-3.00
-1.25
1.25
-1.25
1.25
-1.25
-
7
DPF
(inches)
0.14
1.62
3.04
0.54
3.24
0.31
0.31
0.63
0.63
0.94
11.39
8
1000*
(inches'
2
)
0.54
6.48
23.38
2.16
12.97
1.49
1.49
2.98
2.98
4.46
-
TABLE 1
Results of Virtual Work Analysis of Truss of Figure 1
The values given in column eight for members-3 and 5 were computed as follows:
For 10 cubic inches of material added to member-3:
updated volume = 120 + 10 = 130 inches
3
updated area = 130/120 = 1.0833 inches
2
updated DPF = 1.0*3.04 /1.0833 = 2.8062 inches
change in total displacement = 3.04 - 2.8062 = 0.2338 inches
change in DPF/ change in volume = (0.2338/10)* 1000 = 23.38
For 10 cubic inches of material added to member-5:
updated volume = 240 + 10 = 250 inches
3
updated area = 250/240 = 1.0417 inches
2
updated DPF = 1.0*3.24 / 1.0417 = 3.1103 inches
decrease in total displacement = 3.24 - 3.1103 = 0.1297 inches
change in DPF/ change in volume = (0.1297/10)*1000 = 12.97
It should be noted that the member DPF can be updated without recomputing member forces because this structure
is statically determinate.
As can be seen, member-3 is almost twice as sensitive to a change in volume as is member-5. In the remainder of
this paper the change in member DPF per change in volume is referred to member's sensitivity index (SI).
12-8
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
When the change in volume approaches zero, the sensitivity index is computed as the ratio of the member DPF to
member volume:
(6)
(7)
The final result of equation 6 may also be written as:
which from equation (2) is equal to the volumetric virtual work in the member.
From the above, it is clear that adding material to members with relatively large sensitivity indices is a very effective
way to decrease displacement. This leads to the following observation:
Observation 1:
When adding material to a structure to reduce displacement, the material should be added to the
member(s) with the largest sensitivity index.
For members with very small SIs, volume may be removed from the member with relatively little impact on total
displacement. For example, removing half the volume from member-1 increases its DPF from 0.14 inches to 0.28
inches. This leads to a corollary of Observation 1:
Corollary to Observation 1:
When removing material from a structure to improve economy, the material should be removed
from the member(s) with the smallest sensitivity index.
From the above observations, it is clear that material should be added to members with high SI's, and removed from
members with low SI's. If all members had the same SI, it would appear that there would be no benefit to
redistributing material, because for a given volume of material the structure would be optimum.
To illustrate the fact that a structure with a constant member SIs is optimum, we will redistribute material from one
member to another member in the truss of Figure 3, and in the process, keep the total volume in the two members
constant. This concept is illustrated in Figure 4, which shows how the total DPF for members-3 and 5 changes when
material is moved from member-3 to member-5.
Maintaining the real and virtual axial forces shown in Table 1, the analysis was initiated by distributing 5 percent
of the combined volume of material ( 360 cubic inches) to member-3 and 95 percent of the combined volume to
member-5, and then computing each member's DPF and SI. The new DPFs of each member were added to produce
the total displacement shown on the vertical axis. The value on the horizontal axis represents the ratio of the SI of
member-3 to the SI for member-5. This process was repeated for a 10%-90% volume distribution, a 15%-85%
distribution, and so on until the final distribution was 95% to member-3 and 5% to member-5.
12-9
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
As is observed from Figure 4, the minimum total DPF occurs when the material is distributed such that both
members have identical sensitivity indices. A similar result would occur when distributing a certain volume of
material between any two members. This confirms our hypothesis, and leads to a second observation:
Observation 2:
For a structure with a given volume of material, the deflection in the direction of the virtual force
will be a minimum when the material is distributed such that all members have the same sensitivity
index.
This observation has a corollary which can be stated as follows:
Corollary to Observation 2:
If the material in a structure is distributed such that all members have the same sensitivity index,
moving material from any member to another member will increase the displacement in the
direction of the virtual force.
12-10
Figure 4. Result of Redistribution of Material in Two Members of the Truss
TOTAL DISPLACEMENT CONTRIBUTED BY MEMBERS 3 & 5 (in.)
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
While Observation 2 has the essence of being intuitively obvious, it has been stated as a theorem, and has been
mathematically proven (for trusses) by several engineers working in the area of virtual work optimization (1,19,25).
Instead of showing the proof, the concepts are illustrated using the structure of Figure 3. In this case the total volume
of available material (2080 cubic inches) has been redistributed to produce members with a constant SI. This is done
as a two step process:
Step 1:
Holding the area of member-1 constant determine a new area ( A' ) of all other members such that they have the
same SI as member-1:
Step 2:
Compute the new total volume V , and then adjust all member areas by the factor such that the total volume
is equal to the target volume
The results of this process are shown in Table 2, where it is seen that the total displacement reduces to 8.84 inches.
Each member has a SI of 4.25 (column 8) and the total volume of material is 2080 cubic inches, the same as shown
in Table 1.
For trusses, it can be shown that the relationship between member DPF and SI applies to the whole structure when
the SIs are equal for all members:
(8)
(9)
(10)
(11)
12-11
Equation 11 shows that for an optimum truss, the displacement is proportional to the total volume of material, with
the constant of proportionality being the sensitivity index. Once the member SIs are made equal, it is easy to
reproportion the structure to obtain a minimum displacement for a given of material, or to determine a distribution
of material given a target displacement.
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
1
Member
1
2
3
4
5
6
7
8
9
10
TOTALS
2
Length
L
(inches)
240.00
240.00
120.00
240.00
240.00
200.00
200.00
200.00
200.00
200.00
-
3
Area
A (in
2
)
0.36
1.26
2.44
0,73
1.78
0.61
0.61
0.86
0.86
1.05
-
4
Volume
V(in
3
)
87.29
302.38
292.78
174.58
427.63
121.23
121.23
171.45
171.45
209.98
2080.00
5
Real
Force
R (kips)
7.50
30.00
67.50
-15.00
-45.00
-12.50
12.50
-25.00
25.00
-37.50
-
6
Virtual
Force
Q (kips)
0.75
2.25
3.75
-1.50
-3.00
-1.25
1.25
-1.25
1.25
-1.25
-
7
DPF
(inches)
0.37
1.29
1.24
0.74
1.82
0.52
0.52
0.73
0.73
0.89
8.84
8
1000*
SI
4.25
4.25
4.25
4.25
4.25
4.25
4.25
4.25
4.25
4.25
-
TABLE 2
Results of Optimization of Truss of Figure 1
Equation 12 has been previously presented by Baker (1,2)
(12)
12-12
Instead of using the two-step process outlined above, it can also be shown from equations 3, 7, and 10 that if the
desire is to obtain a minimum weight structure that has a certain target displacement the required cross
sectional area for each member can be obtained in one step as follows:
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Before expanding the presentation to cover rigid frames, four points should be emphasized:
[1] The resizing procedure spelled out above applies to statically determinate structures only. However,
observations 1 and 2 apply to all structures, both determinate and indeterminate (3). The problem with
applying the method to indeterminate structures is that when members are resized, the member forces
change. For indeterminate structures whose member forces are highly sensitive to changes in member
area, iterate by carrying out steps 1 and 2 as described above (or using equation 12) and then reanalyze
the structure to determine new member forces. This process, shown in the flow chart of Figure 2,
should be repeated until the changes in member force with each successive iteration are small.
For determinate or force-insentive indeterminate structures, the fact that updated displacements can be
obtained with out full reanalysis is very beneficial, particularly when the reanalysis is time-consuming.
Using programs like DISPAR, for example, the displacements in large complex structures can be
obtained in real time, meaning that as soon as the member property has been updated, the new
displacement is available. In the same manner, member DPFs and SIs may be imported into spreadsheet
programs, allowing the engineer to perform "what if" calculations on the behavior of the structure being
considered.
[2] It should be noted from Table 2 that some member cross-sectional areas decrease, and some increase
as a result of the resizing process. If any of these areas fall below what would be required for strength,
they should be adjusted upward accordingly. At the end of the resizing process, all stiffness controlled
members would have the same SI, but the strength controlled SIs would be different. The total volume
of the stiffness controlled members could be adjusted, if necessary, to meet the target structure volume
or target drift.
[3] The final areas shown in Table 2 are not very practical. An actual design would use available sizes,
and this would prevent the member SIs from being truly equal. Further, it may be desired to impose
other constraints on the system, such as the requirement that several members be the same unknown
size. In this case, these members would be treated as a group, and the target SI for the group would
be treated the same way as it would be for an individual member.
[4] In Figure 4, note how flat the curve is in the range of SI ratios of 0.1 to 10. For a SI ratio of 0.1 and
10.0, the drift contributed by members-3 and 5 is approximately 40 percent greater than that when the
SIs are equal. When the SI ratios are 0.2 and 5.0, the combined displacement is only 20 percent greater
than optimum. The closer the member is to the optimum SI, the lower the effect on displacement.
This means that when carrying out human-loop optimization, very careful scrutiny should be given to
members with initially low or high SI's. Once all members are within an SI of about 0.5 to 1.5 times
the average SI for all members, very little will be gained by further member resizing.
12-13
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
THE VIRTUAL WORK METHOD AS APPLIED TO RIGID FRAMES
The theoretical development of the virtual work method of behavior identification is easily extended to frame
members. Consider the corner column of the three-dimensional frame of Figure 5. Applying the method of virtual
work, the column's displacement participation factor consists of six components:
(13)
where
and
is the major axis flexural component
is the minor axis flexural component
is the major axis shear component
is the minor axis shear component
is the axial component
is the torsional deformation component.
If the real and virtual member forces are known, the individual components are easily computed. Refer to Figure
6 for details. As mentioned later, the equations in Figure 6 are equally applicable to beams, and the term "L" may
be used to represent either the total span or the clear span length of the member.
12-14
Figure 5. Three-Dimensional Moment Resisting Space Frame
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Figure 6. Components of Deformation in a Typical Frame Element
12-15
REAL FORCE VIRTUAL FORCE DPF
Major Axis Bending
Minor Axis Bending
Major Axis Shear
Minor Axis Shear
Axial
Torsion
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
From the totals under columns 4, 5, and 6 of Table 3, it may be observed that the flexural, shear, and axial
contributions to total displacement are 0.612 inches, 0.108 inches, and 0.0145 inches, respectively. The members
contributing the most to the total displacement are the two interior columns of the first story, with each column
contributing approximately 0.088 inches to the total displacement. As these members also have the greatest
sensitivity index, and since flexural deformation dominates, the total displacement would be most affected by
increasing the moment of inertia of these members. The next most important members are the interior columns of
the second story, closely followed by the exterior columns of the first story.
12-16
As an example, consider the simple planar frame of Figure 7. This three-story, three-bay structure is subjected to
horizontal forces at each story, and a virtual force of 1000 kips at the top level. The columns are W14s, and the
beams are W24s. The structure was analyzed on the computer program SAP90, using centerline dimensions. All
member displacement participation factors and sensitivity indices were computed by the DISPAR postprocessor.
Since this is a planar structure, minor axis flexure, minor axis shear, and torsional contributions to member DPF are
not present. The results of the DISPAR analysis are shown in Table 3, where it may be observed that the total
displacement in the direction of the virtual force is 0.735 inches. (Note that all DPF values in the table include the
1000 multiplier associated with a 1000 kip virtual force). The initial weight of this structure is 23.0 tons.
Figure 7. Simple Moment Resisting Frame
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
1
Member
C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
C11
C12
B1
B2
B3
B4
B5
B6
B7
B8
B9
2
Size
W14x145
W14x145
W14x145
W14x145
W14x120
W14x120
W14x120
W14x120
W1 4x99
W1 4x99
W14x99
W14x99
W24x176
W24x176
W24x176
W24x146
W24x146
W24x146
W24x117
W24x117
W24x117
3
Length
(Inches)
180
180
180
180
150
150
150
150
150
150
150
150
240
240
240
240
240
240
240
240
240
TOTAL STRUCTURE WEIGHT = 23.00 TONS
1000 x Contribution to Total Displacement (inches)
4
Flexural
58.131
76.252
75.591
56.537
17.623
41.411
41 .443
18.171
11 .703
23.382
22.582
10.580
36.525
17.441
36.105
21.704
12.528
20.918
5.432
2.942
4.973
611.975
5
Shear
9.022
12.371
12.263
8.769
3.988
9.288
9.295
4.096
2.581
5.215
5.036
2.324
5.619
2.835
5.559
3.194
1.933
3.086
0.717
0.423
0.660
108.247
6
Axial
2.974
0.205
0.178
2.869
0.725
0.036
0.029
0.690
0.079
0.004
0.003
0.073
-0.138
-0.007
0.047
0.137
0.059
-0.001
4.503
1.746
0.286
14.497
7
Total
70.127
88.828
88.032
68.175
22.336
50.735
50.767
22.956
14.363
28.601
27.621
12.977
42.007
20.629
41.711
25.036
14.520
24.004
10.651
5.111
5.919
734.746
8
1000000 *SI
9.124
11.557
11 .454
8.870
4.218
9.582
9.588
4.336
3.291
6.552
6.328
2.973
3.385
1.634
3.362
2.426
1.407
2.326
1.290
0.619
0.717
TABLES
Virtual Work Analysis of Frame Shown in Figure 7
(Before Optimization)
12-17
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
To illustrate the use of the information in Table 3, we will decrease the total drift to 0.675 inches by increasing the
size of the interior columns of the first story
2
:
The total decrease in drift = 0.735-0.675 = 0.060 inches
Reduction in DPF for each column = 0.060/2 = 0.030 inches
Target Column DPF = 0.088-0.030=0.058
I
NEW
= new moment of inertia per column = I
OLD
(0.088/0.058) = 1.517 I
OLD
For a W14xl45, I
OLD
=1710 inches
4
, so I
NEW
= 1.517x1710=2594 inches
4
Use a W14x211 with I
NEW
=2660 inches
4
.
This change in column size increases the total weight of the structure from 23.00 tons to 23.99 tons. It is interesting
to try the same exercise on the first story exterior beams, which have a DPF of 0.042 inches, and a SI of 1/3 of that
of Column Cl. Assume the drift is to be reduced by resizing Bl and B3:
The total decrease in drift = 0.735-0.675 = 0.060 inches
Reduction in DPF for girder = 0.060/2 = 0.030 inches (assuming the same reduction in each beam)
Target Beam DPF = 0.042-0.030=0.012
I
NEW
= new moment of inertia per beam = I
OLD
(0.042/0.012) = 3.5 I
OLD
For a W24xl76, I
OLD
=5680 inches
4
, so I
NEW
=3.5x5680=19880 inches
4
This requirement can't be met by even the largest W24, which has an I of 19110 inches
4
.
This example emphasizes how important the information in Table 3 is to the design engineer. With the information
at hand, the structure is easily and economically resized. Without it, the designer must guess which members to
resize, and as is evident from the example, wrong guesses can produce very costly designs.
In the above example, it was necessary to add material to the structure to obtain the desired drift. Using the
observation that structures with equal SI's for all members are optimum (1,2), the reduction in total drift can be
accomplished much more effectively than this, and in fact, the total drift can be reduced to less than 0.735 inches
with a significant decrease in total volume.
Unfortunately, the resizing of frame members to achieve a uniform SI is more difficult than it is for trusses. There
are two sources of the difficulty:
[1] Frames are usually statically indeterminate, so changes in member properties affect member forces, and
these in turn affect the computed SI's.
[2] During resizing the DPF and SI associated with the new section property is a nonlinear function of the
area of the section. This requires an iterative solution, which converges when the SI computed using
the new member properties is equal to the target SI.
12-18
This example is similar to problems encountered in seismic retrofit, where a stiffness tolerance must be achieved by modifying existing
members.
2
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
The first source of difficulty can be important for structures that have the property that member forces are highly
sensitive to changes in member sizes. For most moment resisting frames this is not the case, and resizing can
proceed without regard to indeterminacy. The advantage of this procedure is that only one cycle of resizing need be
performed, with a full structural analysis being carried out only twice: once at the beginning of the process, and once
at the end. The first analysis is used to compute the data required for the virtual work analysis, and the last analysis
is required only for the final stress-check. For force-sensitive indeterminate structures several cycles of optimization
must be carried out, and a full structural analysis must be performed at the end of each cycle.
In each cycle of the resizing process the second source of difficulty arises. A remedy for this problem is posed in
step two of the resizing algorithm given below.
Resizing Algorithm for Planar Frames
Resizing of members to attain a target drift with a minimum volume of material is a straightforward process. The
process is iterative for the reasons given above, and for the additional reason that the target drift will not be exactly
attained at the end of the iteration. The steps involved in the process are outlined as follows:
STEP 1: Determine a target sensitivity index for each member. Assuming that the goal is to produce a structure
that attains the target drift with a near minimum volume of material, estimate the target sensitivity index,
as follows:
where is the target drift and is the expected volume of material to be used. The inital volume of
the structure can be used as the first guess for Note that equation 14 attempts to approximately satisfy
equation 11 presented earlier for trusses.
STEP 2: For each member, estimate a new cross sectional area as follows:
the ratio of shear area to total area,
are the inital contributions to the member DPF from flexural, axial, and shear sources of
where r is the radius of gyration, and deformation, repectively. Using
(14)
(15)
If can be assumed that the new area can be estimated from equation 16 as
(16)
12-19
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
12-20
where are constants determined from a least squares fit of the relevant properties of a certain wide
flange shape.
For example, consider the W14 series of sections used in the example. In Figure 8, the variation in with area is
plotted for all section in the W14 series. A similar curve is shown for the variation in 100K with area. For sections
of size W14x90 and above, these relations are essentially linear.
(21)
(20)
and
Now, compare the new SI to the target SI. If the values are within a few percent (say 15%) of each other, save the
new member size and DPF and move to the next member. If the values are significantly different, choose a new
section. For most structures the process should converge in just two or three iterations.
If the iterations are occurring within a constant range of sections (say W 14s), the computations can be carried out
in a continuous fashion instead of using the discrete section sizes available in an AISC database. This is based on
the observation that there exists a linear relationship between major axis radius of gyration squared and cross
sectional area, and between shear area and cross sectional area (3).
(19)
(18)
Where is the original displacement participation factor for the member.
The assumption of equal radius of gyration for the new and old member will be accurate only if the change in area
is small (not more than three or four section increases or decreases within a single shape designation). For sections
in the same depth group, the ratios of shear area to total area are very nearly equal from section to section.
Having a new area, select a steel shape, which has a moment of inertia and a shear area Using these
properties, compute a new DPF and SI for the member:
(17)
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
For the W14 and W24 sections used in this example, the relationships are:
For W14 sections of 90 pounds per foot and greater:
For W24 sections of 104 pounds per foot and greater:
STEP 3: After all new member properties are selected, sum the new DPFs to determine the new displacement. If
this value is significantly different than the target displacement, either go back to Step 1 and iterate, or uniformly
reduce (or increase) all section sizes, one section at a time, until the target drift is reached.
For force-sensitive structures, a new structural analysis should be carried out, and the process should begin again
with STEP 1 and continue until the SI's converge. For force-insensitive structures, a final analysis and stress check
should be performed.
Figure 8. Variation in Member Property with Area for W14 Columns Bent about their Major Axis
12-21
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Using the procedure given above, the 3-story frame of Figure 7 can be optimized very rapidly because it is force-
insensitive to changes in member properties. The results of the process are shown in Table 4, where it can be seen
that (after one iteration) all member Sis are nearly equal. As a result of the resizing, the total drift was reduced from
0.735 inches to 0.720 inches, and the weight reduced from 23.0 tons to only 19.9 tons. This represents a 15%
savings in material over the original design, while the drift was reduced by 2%..
TABLE 4
Virtual Work Analysis of Frame Shown in Figure 7
(After Optimization)
12-22
1
Member
C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
C11
C12
B1
B2
B3
B4
B5
B6
B7
B8
B9
2
Size
W14x176
W1 4x211
W1 4x211
W14x176
W14x109
W14x159
W14x159
W14x109
W14x82
W14x109
W14x109
W14X82
W24X146
W24X103
W24X146
W24x103
W24x68
W24X103
W24x55
W24x55
W24x55
3
Length
(Inches)
180
180
180
180
150
150
150
150
150
150
150
150
240
240
240
240
240
240
240
240
240
TOTAL STRUCTURE WEIGHT = 19.88 TONS
1000 x Contribution to Total Displacement (inches)
4
Flexural
42.401
57.823
56.779
40.483
16.618
33.054
33.000
17.331
13.404
22.945
21 .573
11.070
43.441
26.710
42.470
34.990
22.689
32.718
12.376
8.696
10.665
5
Shear
6.294
9.074
8.908
6.000
3.859
7.835
7.818
4.002
2.261
5.124
4.897
1.863
6.674
3.217
6.526
4.211
2.283
3.948
0.952
0.683
0.826
6
Axial
2.515
0.241
0.199
2.362
0.839
0.055
0.041
0.768
0.088
0.001
0.000
0.077
-0.181
-0.022
0.050
0.236
0.206
0.013
9.748
3.616
0.517
7
Total
51.210
67.138
65.886
48.845
21.316
40.945
40.859
22.101
15.753
28.161
26.471
13.009
49.933
29.906
49.046
39.437
25.178
36.742
23.076
12.994
12.008
8
1000000*SI
5.492
6.016
5.904
5.239
4.411
5.845
5.833
4.604
4.358
5.867
5.515
3.599
4.839
4.112
4.753
5.423
5.219
5.052
5.935
3.342
3.088
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
The final distribution of material for the frame is shown in Figure 9. As may be observed, the resizing process
significantly reduced the size of the beams, and increased the size of the lower story columns. In the lower stories,
material was distributed to the exterior bays, and in the upper story, the interior bay has the larger members. In
a real design, the engineer may wish to have all columns or beams in a particular story the same size. This
constraint can easily be written into the resizing process, although it may be more desirable that the designer make
the final changes. These changes, if not too significant, will have very little effect on the final drift because of the
observation made from Figure 4: all that is required is to have member SI's nearly equal.
Figure 9. Results of Resizing of the 3-Story Frame
12-23
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Application to Three Dimensional Structures
The extension of the resizing procedure to three dimensional structures produces an additional difficulty. Consider,
for example, the structure of Figure 5. In this structure, the real and virtual loads are acting in the direction of the
three-bay frames. When the individual DPFs and SIs are computed for the structure, the values printed for the
members in the two-bay frame (orthgogonal to the load) will be very small because this frame contributes very little
to the lateral resistance in the direction of the loads. In the resizing procedure, these members will be downsized
to the minimum allowed, and the members in the frame parallel to the load would be upsized for stiffness.
When the loads are then applied in the direction of the two-bay frame, the reverse will occur, with the members of
the three-bay frame being downsized and the members of the two-bay frame being increased in size. The problem
of resizing members for orthogonal loading conditions can be avoided if one of the following strategies is employed:
[1] Apply both sets of lateral loads in one load case, with the virtual force being directed in the direction
of the resultant of these forces. This will make all members participate in the resistance. This method
can be used with almost any virtual work postprocessor. Baker has shown the method to be particularly
useful in tube type buildings (1).
[2] Before analysis, identify the members not participating in the lateral load resistance, and eliminate them
from the resizing process. The target SI will be the average of the values for the members actually
participating in the response. This method should only be used for structures with independent
orthogonal lateral load resisting systems. It should not be used for tube type structures, because
members in the frames orthogonal to the load can be very important in the lateral resistance. The
DISPAR program can be set to use this method, where all members with a SI lower than a user
specified portion of the average SI are eliminated from the resizing process.
Inclusion of Beam-Column Joint Deformations
It is well known that deformations occurring in beam-column joints can have a very significant effect on the lateral
stiffness of steel frames (12). In a recent paper by the writer (9), the significance of joint deformations was illustrated
by a series of analyses of 45 different steel framed structures ranging from 10 to 50 stories in height, and with five,
nine, or thirteen bays of 10, 15, or 20 feet in width. The frame with the largest joint deformation component was
the 10-story frame with thirteen 10-foot bays, where joint flexibility contributed 41.3 percent of the total drift. The
frame with the smallest joint deformation component was a 50-story tube with five 10-foot bays, where 15.6 percent
of the total drift was due to joint deformation.
Fortunately, the effect of joint deformation can be easily included in the virtual work analysis. In this method
described below, joint deformations are considered as a seventh component of the member DPF:
The first six terms of equation 22 are computed exactly as shown in Figure 6, except that only those forces occurring
in the clear span of the member are considered. For computational purposes, the joint deformation term, DJ
i
is split
into four components:
(22)
(23)
12-24
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
where
and
is the portion of joint deformation due to joint bending produced by the beam moments
is the portion of joint deformation due to joint bending produced by column moments
is the portion of joint deformation due to joint beam shear.
is the portion of joint deformation due to joint column shear.
The details of the computation of the joint deformation components depend on the idealization of the distribution
of forces within the joint. In the computer program SAP90, for example, the user has the option of considering
a portion of the joint to be "rigid", as shown for a typical beam Figure 10, and for a column in Figure 11. The rigid
portion of the joint extends a distance "z" times the actual eccentricity, where "z" may range from 0.0 (fully flexible
joint) to 1.0 (fully rigid joint). Also shown in the figure are the traditionally assumed distributions of curvature
(M/EI) and shear distortion (V/A*G) in the member. The curvature diagram is linear, with the peak curvature
occurring at the face of the rigid zone of the joint. The shear distortions are constant along the region of the member
bounded by the rigid zones. In the rigid region of the joint, the curvature and the shear distortion are zero.
Using the nomenclature of Figure 10, the beam-column joint deformation components are computed (for either end
of the beam or column) as follows:
(24)
(25)
(26)
(27)
In the above, represent the real and virtual moments throughout the flexible region of the beam portion of
the joint, and and represent the real and virtual shears in the beam portion of the joint. Similarly, the force
terms with Cj subscripts represent force distributions in the flexible portion of the column. It should be noted that
the functions in the integrals of equations 24 through 26 are straight lines for moment, and constants for shear.
The quantities in equations 24 and 26 are applied to each end of the beam, and are added to the components
computed for the clearspan to obtain the total member DPF. For the columns, equations 25 and 27 are applied to
each end, and added to the clearspan terms to obtain the total member DPF.
12-25
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Figure 10. Modelling of Beams to Include Beam-Column Joint Deformation
Figure 11. Modeling of Columns to Include Beam-Column Joint Deformations
12-26
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
JOINT MODEL
SAP90 z=0.0
SAP90 z=0.25
SAP90 z=0.50
SAP90 z=0.75
SAP90 z=1.0
"TRUE" FLEXIBLE
JOINT
CLEAR SPAN
DISPLACEMENT
(inches)
0.577
0.577
0.577
0.577
0.577
0.577
JOINT
DISPLACEMENT
(inches)
0.143
0.105
0.068
0.033
0.000
0.264
TOTAL
DISPLACEMENT
(inches)
0.720
0.682
0.645
0.610
0.577
0.841
COMMENT
FULLY FLEXIBLE
-
-
-
FULLY RIGID
NO BEAM FLANGE
CONTINUITY
PLATES
Table 5
Effect of Joint Deformations on Frame Deflections
As may be observed from Table 5, the inclusion of joint deformation has a very large effect on the total drift of the
structure. The SAP90 Users' Manual suggests that for most structures a "z" value of 0.5 is appropriate.
Unfortunately there is little basis for this recommendation, since true joint modeling shows that the total drift for
this structure is greater than the drift computed with centerline analysis. The basis for this result, which is shown
in the last row of Table 5, is now developed.
Figure 12 shows a blow-up of a typical interior subassemblage of a planar frame, and a further expansion of the
beam-column joint itself. The forces on the joint include the direct shears Vc and Vb transmitted by the beams and
columns, and an additional set of forces designated as either Mc/dc or Mb/db. These latter forces are
approximations of the column flange forces and the beam flange forces entering the joint. These forces have two
pronounced effects:
[1] The column flange forces react at the face of the column, and cause a change in shape of the beam
bending moment and curvature distribution within the joint. The beam flange forces have a similar
effect on the columns. The resulting curvature diagrams are approximated in Figures 10 (E) and 11(E)
3
.
[2] There is a large increase in the shear forces and shear distortion occurring within the panel of the beam-
column joint. This effect is shown in Figures 10 (F) and 11(F).
While the digrams show curvature (moment) passing through zero at the center of the joint, there are small momenta at the joint center.
These moments are negligible in the computation of joint flexural deformation.
12-27
Using the DISPAR program, the above equations were applied to the frame of Figure 7. In the analysis, the rigid
end zone factor z was varied between 0.0 and 1.0, with results as shown in Table 5. In the development of the table,
it should be noted that each analysis was based on the member forces computed with z=0.0. Since the structure is
force-insensitive, recalculation of DPFs without reanalysis is very accurate.
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Figure 12. Forces in Beam-Column Joints
If the revised curvature and shear distortions are used in lieu of those shown in parts C and D of Figures 10 and
11, equations 24 and 26 can be applied as shown (with z=0.0), as well as can either equation 25 or 27. In this case,
only one shear equation should be considered because the beam and column shear distortions occur over the same
volume of material. It is suggested that equation 27 be used because the shear area in the joint is typically supplied
by column web material. When applying equation 27, the shear at each end of the joint is computed from Figure
12. The horizontal joint shear force is:
The vertical joint shear force is:
From equation 28 and 29, it can be shown that the column joint shear can be represented as:
12-28
(28)
(29)
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
When "true" joint deformations are used in the analysis of the 3-story frame, the drift increases to 0.841 inches, as
shown in the last row of Table 5. This value is larger than the clearspan drift of 0.720 inches.
The "true" joint model described above has been verified by finite element analysis of more than thrity beam-column
subassemblages (12). This analysis showed that for joints without beam flange continuity plates, the error was
generally less than five percent. When continuity plates are used, the true joint model overestimates joint
deformation by approximately ten percent.
(30)
SPECIAL TOPICS IN THE USE OF THE VIRTUAL WORK METHOD
Other Virtual Force Patterns
In the previous examples, it was assumed that the objective of the virtual work analysis was to optimize the structure
for total drift, so the virtual force was placed in the direction of the roof displacement. There are, or course, other
critical deflections in structures. Fortunately, the virtual force method is applicable to these cases as well. Four
examples of structures where the virtual force is not placed at the roof are shown in Figure 13.
Figure 13a illustrates a situation where a soft story is known to exist. It is desired to increase the stiffness of the
story in the most effective way possible. The virtual force applied at the first story will produce DPFs and SIs that
indicate how each member contributes to the first story drift.
In Figure 13b, a minimum weight structure that vibrates at a known target frequency of vibration is desired.
Henige (16) has shown that if both the real and the virtual loads are set equal to the inertial force profile under the
assumed mode shape, the DPFs so obtained will indicate how much each member contributes to the frequency. This
type of problem often occurs when trying to design to a specified human comfort related acceleration tolerance.
While lateral displacements are often the controlling deflection for buildings, structures such as the microwave
antenna tower shown in Figure 13c are sensitive to rotation about a horizontal vector orthogonal to the "line of sight"
of the antenna. Limits on such motion, say plus or minus 0.5 degrees, are established to make sure that the antenna
does not send its signal astray during high winds. Since the critical displacement is rotation, the virtual force is a
unit moment applied the direction of the controlling rotation. In this case, the individual DPFs show how much each
member contributes to the rotation, and the corresponding SIs can be used to optimize the structure.
For buildings with unusual plan configurations, or structures with non-coincident centers of mass and stiffness,
rotations can develop about a vertical axis when lateral loads are applied. If these rotations are excessive, drift and
accelerations at the perimeter frames can be much higher than those computed at the center of mass. To determine
how much each member contributes to the rotation, place a virtual moment about the vertical axis, as shown in
Figure 13d. Using the DPFs and SIs obtained, the members may be resized to reduce, or even eliminate the rotation.
12-29
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
Figure 13. Examples of Other Useful Virtual Force Patterns
Negative DPFs
Whenever the real member force and the virtual member force are of opposite sign, the computed DPF and SI will
be negative. This indicates that, for this member, increasing the members size will increase the displacement in the
direction of the virtual force. In most cases, members with large negative DPFs should be removed from the
structural system. However, before doing so, the designer should be sure that the negative DPFs are not simply a
manifestation of localized behavior associated with the placement of real and virtual forces.
NEEDS FOR FUTURE RESEARCH
It is the writer's opinion that the use of the computer programs based on the virtual force method of optimization
and behavior identification will eventually be as common a tool in design offices as are today's structural analysis,
structural design, and CAD programs. However, this will occur only if the techniques are developed into programs
that give the designer complete control of how the information is used.
The evolution of the virtual work method has occurred in three phases, as listed below. The first phase is almost
fully developed, but most of the second and third phases are in their infancy, and are candidates for future research.
[PHASE 1] Printing and plotting of member DPFs and SIs: This information is already supplied by most virtual
work postprocessors.
12-30
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
[PHASE 2] Interactive resizing of members: Here, the designer may instantly see the effect of a change in member
size on the performance/behavior of the structure. In the short term, this can be accomplished by spread sheet
calculations, but in the long term, this should be done through real-time interaction with computer aided drafting
programs. The ability to interact with spread sheets is currently functional in several virtual work postprocessors,
but graphic interaction has not yet been developed.
[PHASE 3] Automatic resizing: This capability exists in many of the currently existing postprocessors, but has not
yet developed to the extent that resizing accounts for multiple real/virtual load patterns, member size constraints, and
member groups. Automatic resizing with multiple constraints does exist in most of the formal optimization programs,
but these programs have the limitation that no control is given to the designer in interpreting or using the internal
DPF and SI information.
The final aspect of the virtual force method that has yet to be applied is in structural engineering education. In
undergraduate structural analysis classes, virtual work is taught as a means to an end ... computing deflections. As
seen in this paper, the method is much more powerful than this, because it has the capability to open a window to
the understanding of structural behavior.
12-31
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
REFERENCES
(1) Baker, W.F., "Sizing Techniques for Lateral Systems in Multi-Story Steel Buildings", Proceedings of the
Fourth World Congress, Tall Buildings: 2000 and Beyond, Council on Tall Buildings and Urban Habitat,
Hong Kong, 1990.
(2) Baker, W. F., "Stiffness Optimization Methods for Lateral Systems of Buildings: A Theoretical Basis",
Proceedings of the Tenth Conference on Electronic Computation, ASCE, Indianapolis, Indiana, 1991.
(3) Balling, R. , "A Close Look at Optimization Via Virtual Work", Proceedings of the 1992 Structures
Congress, ASCE, San Antonio, Texas, 1992.
(4) Brazil, A., Joseph, L., and Scarangello, T., "Efficiency Coefficient Optimization in a Design Office",
Proceedings of the 1992 Structures Congress, ASCE, San Antonio, Texas, 1992.
(5) Cameron, G.E., Xu, L., and Grierson, D.E., "Discrete Optimal Design of 3D Frameworks, Proceedings of
the Tenth Conference on Electronic Computation, ASCE, Indianapolis, Indiana, 1991.
(6) Chan, C., and Grierson, D.E., "Optimality Conditions for Lateral-Load Resisting Tall Steel Buildings
Frameworks Subject to Multiple Drift Constraints", Proceedings of the 1992 Structures Congress, ASCE, San
Antonio, Texas, 1992.
(7) Charney, F.A., "DISPAR for ETABS, User's Manual", Advanced Structural Concepts, Denver, Colorado,
1990.
(8) Charney, F.A., "DISPAR for SAP90, User's Manual", Advanced Structural Concepts, Denver, Colorado,
1990.
(9) Charney, F.A., "Sources of Elastic Deformations in Laterally Loaded Steel Frame and Tube Structures",
Proceedings of the Fourth World Congress, Tall Buildings: 2000 and Beyond, Council on Tall Buildings
and Urban Habitat, Hong Kong, 1990.
(10) Charney, F.A., "The Use of Displacement Participation Factors in the Optimization of Wind Drift Controlled
Buildings", Proceeding of the Second Conference on Tall Buildings in Seismic Regions, Council on Tall
Buildings and Urban Habitat, Los Angeles, 1991.
(11) Charney, F.A., "The Use of Displacement Participation Factors in the Optimization of Wind Sensitive
Buildings", Proceeding of the 1991 Structures Congress, ASCE, Indianapolis, Indiana,1991.
(12) Charney, F.A., and Johnson, R., "The Effect of Panel Zone Deformations on the Drift of Steel Framed
Structures," Presented at the ASCE Structures Congress, New Orleans, September 1986.
(13) Forrest-Brown, G., and Samali, B., "Practical Optimisation of Framed Structures Using Virtual Work
Principles", Proceedings of the Fourth World Congress, Tall Buildings: 2000 and Beyond, Council on Tall
Buildings and Urban Habitat, Hong Kong, 1990.
(14) Grierson, D.E., and Cameron, G.E., "Microcomputer-Based Optimization of Steel Structures in Professional
Practice", Microcomputers in Civil Engineering, Elsevier Science Publishing, 1989.
12-32
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.
(15) Grierson, D.E., and Lee, W.H. "Optimal Synthesis of Steel Frameworks Using Standard Sections", ASCE
Journal of Structural Mechanics, Volume 13, Number 3, 1984.
(16) Henige, R. A., "Structural Optimization to Limit Natural Period", Proceedings of the Tenth Conference on
Electronic Computation, ASCE, Indianapolis, Indiana, 1991.
(17) lyengar, H., and Sinn, R., "Structural Approximations in Multi-Story Buildings", Proceedings of the 1991
Structures Congress, ASCE, Indianapolis, Indiana, 1991.
(18) McNamara, R.J., "Simplified Insights into the Behavior of Complex Structures", Proceedings of the 1992
Structures Congress, ASCE, San Antonio, Texas, 1992.
(19) Thornton, C. H., Joseph, L., and Scarangello, T.Z., "Approximate Preliminary Methods for Design of Lateral
Load Resisting Systems in High Rise Buildings", Proceedings of the 1991 Structures Congress, ASCE,
Indianapolis, Indiana, 1991.
(20) Velivasakis, E.E., and DeScenza, R., "Design Optimization of Lateral Load Resisting Frameworks",
Proceedings of the Eight Conference on Electronic Computation, ASCE, Houston, Texas, 1983.
(21) Wada, A., "Drift Control Method for Structural Design of Tall Buildings", Proceeding of the Second
Conference on Tall Buildings in Seismic Regions, Council on Tall Buildings and Urban Habitat, Los Angeles,
1991.
(22) Wada, A., "Structure Improvement Method Focusing on Lateral Drift", Proceedings of the 1992 Structures
Congress, ASCE, San Antonio, Texas, 1992.
12-33
2003 by American Institute of Steel Construction, Inc. All rights reserved.
This publication or any part thereof must not be reproduced in any form without permission of the publisher.

Potrebbero piacerti anche