Sunteți pe pagina 1din 242

Extending performance-based design methods by applying

structural engineering design patterns


by
John-Michael Wong

B.S. (University of California, Berkeley) 2003


M.S. (University of California, Berkeley) 2004

A dissertation submitted in partial satisfaction of the


requirements for the degree of
Doctor of Philosophy
in
EngineeringCivil and Environmental Engineering
in the
Graduate Division
of the
University of California, Berkeley

Committee in charge:
Professor Bozidar Stojadinovic, Chair
Professor Stephen A. Mahin
Professor James W. Demmel
Fall 2008

UMI Number: 3353354

INFORMATION TO USERS

The quality of this reproduction is dependent upon the quality of the copy
submitted. Broken or indistinct print, colored or poor quality illustrations and
photographs, print bleed-through, substandard margins, and improper
alignment can adversely affect reproduction.
In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if unauthorized
copyright material had to be removed, a note will indicate the deletion.

UMI
UMI Microform 3353354
Copyright 2009 by ProQuest LLC.
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.
ProQuest LLC
789 E. Eisenhower Parkway
PO Box 1346
Ann Arbor, Ml 48106-1346

Extending performance-based design methods by applying


structural engineering design patterns
2008
by
John-Michael Wong

1
Abstract
Extending performance-based design methods by applying
structural engineering design patterns
by
John-Michael Wong
Doctor of Philosophy in EngineeringCivil and Environmental Engineering
University of California, Berkeley
Professor Bozidar Stojadinovic, Chair

Performance-based design methods are powerful tools for designing structures that are
safe, reliable, and predictable in an environment with rare but potentially catastrophic hazards, such as earthquakes. The design methods used today are made for the just-built state
of the structure, and place relatively little emphasis on other stages in the structure's life cycle. Current design methods need extensions that specifically consider performance across
the entire life cycle of the structure, for example for the construction stage or the stage
of post-earthquake repairs. Such considerations are challenging during the early stages of
design because the precision necessary for analyzing structural performance during, for
example, construction and post-earthquake repair stages is typically unavailable until too
late after design decisions are made. Therefore, the task of extending performance-based
design requires an analysis of design information, analysis methods, and decision making.

2
Those needs are fulfilled in this thesis by investigating structural engineering analysis and
decision making methods using the design pattern method, and by defining the qualities
and types of design information.
Patterns describe procedures and practices for solving recurring problems in a general
way, which allows comparison and understanding of the tradeoffs involved in their use.
The definition of structural engineering design information categories reveals what information is needed to perform certain types of analysis for construction and repair, and how
to communicate it at the appropriate level of detail. This description explains the different
types of information needed for conceptual design and for construction. The discussion
of analysis and decision methods using patterns identifies unnecessarily premature decisions in existing structural design methods, and locations where mechanics models alone
are insufficient for understanding structural performance. The pattern definitions also allow
concise expression of design intent within structural engineering problem solving methods.
The modifications to performance-based design methods, developed through examples
using the pattern framework and presented in this thesis, incorporate performance considerations for all phases of the structure life cycle. These recommendations include: considering additional analysis patterns during design, using decision patterns that account
for performance in all life cycle phases, and implementing set-based design methods for
evaluating multiple design options.

For from Him and through Him and to Him are all things.
To Him be the glory forever.
Amen.
Apostle Paul (Romans 11:36 NASB)

ii

Contents
List of Figures

List of Tables

vi

List of Symbols and Acronyms

vii

Introduction
1.1 What is this thesis about?
1.2 Why might this thesis be worth reading?
1.3 How can this thesis be read most effectively?
1.3.1 Other work
1.3.2 Chapter descriptions
1.3.3 Patterns

1
1
3
5
6
9
11

Performance-centric design
2.1 Concept phase
2.2 Design phase
2.3 Construction phase
2.4 Operation phase
2.5 Repair phase
2.6 Knowledge reuse
2.7 Summary

16
18
19
22
23
24
24
26

Information in structural design


3.1 Content of information
3.1.1 Material
3.1.2 Functional
3.1.3 Spatial
3.1.4 Process
3.1.5 Abstract

27
30
30
30
30
31
31

iii
3.2

3.3

3.4
4

Quality of information
3.2.1 Detail
3.2.2 Ambiguity
3.2.3 Maturity
3.2.4 Precision
3.2.5 Accuracy
3.2.6 Grouping: agglomeration and aggregation
Importance of content and quality
3.3.1 Mechanics models
3.3.2 Communicating quality of information
Summary

Design communication patterns


4.1 Communication fundamentals and hermeneutics
4.2 Patterns for communication
4.2.1 Narrative pattern
4.2.2 Sketch pattern
4.2.3 Annotation pattern
4.2.4 Spreadsheet pattern
4.2.5 Mechanics Model pattern
4.2.6 Rebar Placing Drawings pattern
4.2.7 Database pattern
4.2.8 Schedule pattern
4.3 Application to highway bridge loss estimation
4.4 Application to building information modeling
4.5 Summary

5 Analysis and decision patterns


5.1 Patterns for analysis
5.1.1 Abstraction pattern
5.1.2 Aggregation pattern
5.1.3 Agglomeration pattern
5.1.4 Updating pattern
5.1.5 Engineering Performance pattern
5.1.6 Social Performance pattern
5.2 Patterns for decision making
5.2.1 Pick Any pattern
5.2.2 Minimum Weight pattern
5.2.3 Minimum Cost pattern
5.2.4 Optimization pattern
5.2.5 Prescription pattern
5.2.6 Set Exploration pattern .

32
32
32
33
36
37
37
39
39
42
44
45
46
53
53
59
64
68
73
76
83
87
91
102
110
111
113
113
118
122
125
129
132
136
136
140
143
146
149
153

iv
5.3
5.4

Metadata for design process and design product


Examples of design and decision patterns
5.4.1 Scope of examples
5.4.2 Reinforced concrete beam design
5.4.3 Preliminary frame design
5.4.4 FEMA 350 steel moment connections
Summary

157
159
161
161
168
174
180

Extending performance-based design


6.1 Extension 1: Additional analysis for decision support
6.2 Extension 2: Change decision patterns
6.3 Extension 3: Use set-based decision patterns
6.4 Summary

181
182
188
192
197

Conclusion
7.1 What I have done here
7.2 Future work
7.2.1 Expanded pattern catalog
7.2.2 Communication patterns with BIM
7.2.3 Teaching and training engineers
7.2.4 Design with high ambiguity and unclear mechanics
7.2.5 Reliability assessment during design
7.2.6 Development of computer tools for supporting set-based design . .

198
198
200
200
201
202
204
205
206

5.5

A Pattern summary
A.l Communication patterns
A.2 Analysis patterns
A.3 Decision patterns

223
223
224
225

List of Figures
Fig. 2.1
Fig. 3.1
Fig. 4.1
Fig. 4.2
Fig. 4.3
Fig. 4.4
Fig. 4.5
Fig. 4.6
Fig. 4.7
Fig. 4.8
Fig. 4.9
Fig. 5.1
Fig. 5.2
Fig. 5.3
Fig. 5.4
Fig. 5.5
Fig. 5.6
Fig. 5.7
Fig. 5.8
Fig. 6.1
Fig. 6.2
Fig. 6.3
Fig. 6.4
Fig. 6.5
Fig. 6.6
Fig. 6.7
Fig. 7.1

Project phases
Design variables for structure configuration
Narratives describe structural design at a functional level
Sketch example showing preliminary structural system
Sketch example showing preliminary floor system
Annotations enable proper interpretation
Unit repair cost spreadsheet
Mechanics model example
Placing drawings example
Relational data model for bridge performance database
Schedule for bridge repair scenario task
Updating pattern
Social performance
Pick Any pattern
Minimum weight design aid
Set exploration example for beam-column rebar
Inelastic behavior of frames with hinges in beam span
Location of plastic hinge formation
Reduced beam section connection detail
Additional analysis
Considering construction phase performance
Labor productivity rates for different bar sizes
Changing decision patterns
Using set-based decision patterns
Required steel area based on mechanics model
Bar layout options for beam and column
Set-based design prototype tool

17
28
55
62
63
66
71
75
79
86
89
127
134
139
142
155
176
178
179
183
185
186
191
193
195
196
207

VI

List of Tables
Table 1.1
Table 3.1
Table 3.2
Table 4.1
Table 4.2
Table 5.1
Table 5.2
Table 5.3
Table 5.4
Table 6.1
Table 6.2

Pattern template
Information content types answer different questions
Information quality types answer different questions
Qualities of communication patterns
Loss estimation procedure for reinforced concrete bridges
Design procedure for known b and h
Target drift and yield mechanism method
Connection design procedure based on FEMA 350 basic design approach
Steel moment-resisting connection design procedure
Original procedure for single column bent design
Improved procedure for single column bent design

14
29
29
49
93
163
168
175
178
187
189

Vll

List of Symbols and Acronyms


The following symbols and acronyms are used in this dissertation:
a
ab
Gtcl

A
Aeff

As
ACI
ASTM

=
=
=
=
=
=
=
=

ATC =
b =

K
BDA
BDD
BIM
Ce
CAD
CBC
CRSI
d
Dc
Ds
DL
DM
DS
DV
E
Ec

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

depth of equivalent rectangular compression stress block


depth of rectangular stress block at balanced strain conditions
depth of stress block at the tension-controlled limit
cross-sectional area
effective area
cross-sectional area of steel
American Concrete Institute
ASTM International, formerly American Society for Testing and Materials
Applied Technology Council
beam width, or mode shape factor
width of web
Caltrans Bridge Design Aids
Caltrans Bridge Design Details
building information modeling
wind exposure factor for gust and height, or seismic coefficient
computer-aided design
California Building Code
Concrete Reinforcing Steel Institute
depth from top of beam to steel centroid
column cross section depth
bent cap depth
dead load
damage model
damage state
decision variable
elastic modulus
elastic modulus of concrete

Vlll

Es
EDP
FEMA
f'c
Fi
FiU
Fy, fy
h
H
/
IM
jd

=
=
=
=
=
=
=
=
=
=
=
=

LL
LLRCAT
LRFD
M
Mc
Mf
Mn
Mp
Mpb
Mpt,r
Mpc
Mpr
Mu
My
Myf
n
PBEE
PEER
PG
PPC
RBS
RCR
Sf,
s
SDC
T,T\
Vf,
Vp
W

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

elastic modulus of steel


engineering demand parameter
Federal Emergency Management Agency
unconfined concrete compressive strength
design forces at the /th story
updated design forces at the nth story
steel yield strength
beam height
building height
moment of inertia for bending
intensity measure
distance between compression and tension resultants inside a cross section
live load
local linearization repair cost and time methodology
load and resistance factor design
moment
moment at centerline
moment at column face
modal mass, or nominal moment strength
plastic moment
plastic moment of beam
reference beam plastic moment
plastic moment of column
probable plastic moment
ultimate moment
yield moment
yield moment at column face
modulus ratio
performance-based earthquake engineering
Pacific Earthquake Engineering Research Center
performance group
percent planned complete
reduced beam section
repair cost ratio
plastic hinge location
spatial distribution of external forces
Caltrans Seismic Design Criteria
fundamental period
base shear
shear at plastic hinge
building weight

wc
Zx

=
=

Pi
Pi

=
=

&CU

e
e,

%>

=
=

&

P
Pb

(0

unit weight of concrete


plastic modulus, plastic modulus about the strong
rectangular stress block coefficient
shear proportioning factor
modal coefficient
shape proportioning modification factor
compressive strain at crushing of concrete
curvature
target drift
yield drift
shear proportioning factor
displacement ductility
overstrength factor at /th story
reinforcement ratio
balanced reinforcement ratio
resistance factor
mode shape of nth mode
reinforcing index

Acknowledgments
I am thankful to the Lord Jesus Christ for giving me new life, the privilege of learning
how to learn, and opportunities to serve others while studying at the University of California, Berkeley.
I am particularly thankful for my advisor, Professor Bozidar Stojadinovic. He provided
me the freedom, resources, and guidance to pursue a broad spectrum of research topics
within structural engineering and beyond. His flexibility and consistent encouragement
helped me navigate the course of my studies since my junior year of college. Boza is
among the best of the best and I am privileged to have worked with him.
The research on reinforced concrete bridges would have been impossible without the
teamwork of Professor Kevin R. Mackie. He helped me understand the ongoing work
within the Pacific Engineering Earthquake Research (PEER) Center, and his insightful advice and active work ethic made our research progress at record speed. I am thankful for his
assistance in helping develop my technical communication skills and fostering an attitude
of academic diligence.
The completion of my doctoral studies would not have been possible without the camaraderie of my fellow students. My long time officemates, Ady Aviram and Jeff Hunt, have
been a consistent source of academic sharpening and motivation. Kristen Parrish helped
me discover the world of design methodology and project management, and has been a
consistent source of cheerfulness. Matthew Dryden, Hong Kim, and Seongeon Jeong have
provided me with enduring fellowship throughout my doctoral studies and I very much

xi

look forward to co-laboring with them in the time to come.


I am thankful for the guidance and encouragement of Professors James W. Demmel,
Steven D. Glaser, Stephen A. Mahin, and Khalid M. Mosalam. Their service on my qualifying examination and dissertation committees completed the shape of this dissertation.
A special note of acknowledgment is due to Professors Iris D. Tommelein and H. Glenn
Ballard, for their gracious support and instruction in many academic areas beyond structural
engineering. Much of this dissertation is a product of the many synergies developed through
my interactions with them and the research initiatives within the P2SL laboratory.
My family has been a consistent source of strength, and has contributed to the completion of this dissertation in ways that 1 am unable to fully articulate. I am thankful for my
mom's sacrificial love and labor towards me. My brothers and sisters in the church have
spurred me on to love and good deeds and helped me maintain a proper balance among all
of my numerous roles and responsibilities. I am also particularly thankful for my fiancee
Clara and her encouragement, prayers, and flexible support during the final stages of my
doctoral studies.

Chapter 1
Introduction
This chapter answers the following questions: What is this thesis about? Why might it
be worth reading? How can it be read most easily and effectively?

1.1 What is this thesis about?


Current performance-based design methods focus primarily on structural behavior during normal service and in a single earthquake event, but with relatively little emphasis on
other stages in the structure life cycle. This thesis develops new methods for specifically including constructibility and reparability as performance measures for consideration during
design.
In order to accomplish this, I need to investigate basic design questions such as: When
are mechanics models sufficient, or insufficient, for making design decisions? If the mechanics models are insufficient, then what additional information is needed for making

decisions? What design prescriptions exist that must be followed, and why? What nonengineering factors drive design decisions, and where should they come into the design
process? If structural engineering-based models lack content needed to make design decisions, could a new model or methodology connect decisions back to mechanics fundamentals?
These types of questions require fresh thinking about the nature of the design process
and the intentions behind current structural engineering design procedures. Basic terms
are defined which describe the categories and qualities of design information. The "pattern" paradigm is employed to describe methods of communicating preliminary design
information, and the basic types of analysis and decision making that occur inside design
procedures. This provides a common platform for critiquing current design methods and
for understanding where the recommended design extensions fit into a design process.
This thesis describes the differences between quantities that can and cannot be determined using structural engineering methods. Conventional mechanics models for structural
analysis are unable to describe a fully complete design problem because those models are
simply unable to consider non-mechanical issues. Parameters such as cost, construction
time, and material availability are essential for the design process, yet are difficult or impossible to characterize within a mechanics-based model without using prescriptive design
constraints . The PEER PBEE methodology (Cornell and Krawinkler, 2000; Moehle et al.,
2005) is one example of attempting to bridge the gap between mechanics-based analysis
results ("engineering demand parameters") and non-mechanical factors such as cost and

time ("decision variables").


This description is important for achieving true performance-based design because
"performance" need not refer solely to structural response criteria (e.g. inter-story drift,
maximum moment, peak strains), but should also involve other factors such as cost, construction time, reduced uncertainty, extent of earthquake damage, and anticipated downtime
and repair. In this expanded sense of "performance", correct performance-centric thinking
must be applied for all project phases, not just for the as-built structure.
This thesis applies the set-based design method for incorporating these additional measures of performance into structural engineering design. Set-based design evaluates many
feasible design solutions for as long as possible in order to gain the advantage of acquiring
more information. Design decisions are postponed until the last responsible moment so
that viable alternatives may be considered for their influence on the entire project.

1.2

Why might this thesis be worth reading?

This thesis is worth reading because it explains a method for including non-mechanics
based criteria into the design process, such as construction performance. Construction performance is an important design consideration, but is often based on factors outside the mechanics model. This thesis also considers structural performance beyond the initial as-built
state by including possible post-earthquake repairs and constructibility issues. Earthquake
repairs are an appropriate consideration in high seismic hazard areas like California, and

especially for critical infrastructure projects such as highway bridges.


The main intellectual contributions of this thesis include:
1. Identification of unnecessarily premature decisions in existing structural engineering design methods. This is accomplished by showing where and how decisions
are made using pattern descriptions. Describing decision-making methods that are
tacitly embedded in textbook design procedures reveals shortcomings that can be
easily improved by substituting a better decision making pattern in its place.
2. Identification of decision points in existing structural engineering design methods that assume the information from a mechanics model is sufficient, when in
fact the controlling variables may not be based on a mechanics model. Other
information not based on mechanics models such as cost, constructibility, reparability, time, and material availability are also important for making sound structural
engineering design decisions. This is examined by analyzing patterns for structural
engineering analysis and decision making.
3. Definition of different categories of structural engineering design information.
These categories are needed for expressing design intent and selecting the appropriate
methods of communicating design information.
4. Explanation of patterns for design communication, analysis, and decision making. Different patterns of design communication are explained with their suitability
for communicating the different types of structural engineering design information.
The patterns of analysis and decision making separate evaluation from design selec-

tion and reveal areas where more analysis can help inform better decisions.
5. Definition of metadata for analysis and decision making patterns, which are
useful for expressing design intent in structural engineering problem solving
algorithms. Knowing only the final design information, but without the underlying
information used to arrive at those decisions, can cause rework when the final design
turns out to be infeasible later in the design or construction process. Documenting
the reasons why decisions were made and why alternatives were discarded would
reduce this rework and promote learning.
6. Explanation that set-based design naturally fits with a proper understanding
of structural design information and the results from mechanics models used in
design. Mechanics-based design criteria are typically unable to indicate only one optimal design, but often leave several choices that can fulfill the requirements. For example, designing the required area of reinforcing steel for reinforced concrete beam
flexure leaves several possible options for bar size and configuration. This leaves
room for a set of options to be evaluated.

1.3

How can this thesis be read most effectively?

This thesis can be read most effectively by understanding its context in light of my
other works, by understanding the organization of the chapters, and understanding the use
of patterns to describe methods of design.

1.3.1

Other work

This thesis is best understood in the context of my other works, which generally fall into
three categories: set-based design, highway bridge loss modeling for earthquake hazards,
and structural state data and metadata. The work on set-based design complements the
understanding of the Explore Set pattern and the advancements proposed in Chapter 6.
Chapter 4 uses an example of damage scenarios and performance groups from my work
on reinforced concrete bridge loss modeling. This thesis discusses the creation of damage
scenarios from a more general perspective than the detailed examples in Mackie et al.
(2008a), and suggests how bridge loss information might fit into a bridge design process.

Set-based design
Set-based design is a methodology for maintaining feasible design solutions for longer
in the design process than is otherwise affordable using point-based design, for the purpose
of obtaining input from several project participants simultaneously and early on. Set-based
design supports communication between these parties so that they can develop a more
globally satisfactory design solution while reducing rework.
An introduction to set-based design for structural engineering and reinforced concrete
design is presented in Parrish et al. (2007). That paper describes the general concept of setbased design from manufacturing and applies the concept to structural engineering. Setbased design is also described in its connection to lean construction. Parrish et al. (2008a)
explains a particular examples of set-based design for a new hospital in San Francisco,

7
California. That case study example describes the selection of the lateral force resisting
structural system, beam orientation, wall penetrations, and structure skin. Those examples
show an example of the set-based design method and structural engineering decisions that
were made based on construction performance. Parrish et al. (2008b) develops the concept
of "value propositions" which provide information about how different project stakeholders
rank different design options. An example is given for performance measures related to
rebar fabrication and placing time with different sizes and bending types of rebar.
The work on set-based design explains concepts behind the Explore Set pattern (5.2.6)
in more complete detail, and elaborates on the rationale for postponing commitment from
the lean construction perspective. The case studies present real examples of how set-based
design methods can be used in structural engineering practice at the level of preliminary design and at the detail needed for construction. The work on value propositions and communication gives several examples of analysis patterns for Social Performance (5.1.6). These
provide a glimpse into the performance measures that are important for the construction
phase and repair phase of a project.

Highway bridge loss modeling


Mackie et al. (2008a) contains the entire description of a method for performancebased evaluation of reinforced concrete highway bridges including post-earthquake repair
costs and time. This introduces the local linearization repair cost and time methodology
(LLRCAT) for computing probabilistic repair cost and repair times to bridge components

8
for varying degrees of damage. This method retains the simplicity of automated, closedform solutions and creates a well-behaved model that links damage to repair quantity. The
LLRCAT method is summarized in Mackie et al. (2008b,c), and the data needed to support
the methodology are summarized in Wong et al. (2008).
The work on highway bridge loss modeling demonstrates and addresses the challenge of
design information availability in extending performance-based design into the repair and
operation phases. The level of precision needed to detail repairs is not yet available early in
the project during preliminary design. But those early stages are exactly where information
about construction and repair are the most useful, because design changes are the most
feasible and effective earlier rather than later. This challenge is overcome by creating
damage scenarios that provide just enough information for estimating repair quantity, cost,
and time, but with only limited information about the structure. This procedure is explained
in Chapter 4 using communication patterns.

Structural state data and metadata


Stojadinovic et al. (2004) describes a framework for integrating structural state data
from sensors. This framework defines metadata standards for sensor information, included
web-based visualization tools, and was demonstrated using wireless accelerometer data
from a shaking table test of a reinforced concrete bridge column. Wong and Stojadinovic
(2005) summarizes the specific work on metadata descriptions and the web interface, and
Wong et al. (2005) summarizes the work on wireless sensors, data acquisition, and data in-

9
gestion. Wong et al. (2006) proposes a method for using sensor data to improve estimates
on bridge performance for post-earthquake repair cost and repair time evaluation. Wong
et al. (2007) describes state data and metadata for the design process and compares the
structure of point-based and set-based design methods. Discrete event simulations demonstrate the types of communication needed for set-based design.
The idea of using structural state data to update our understanding of a structure's performance is a parallel concept to the Updating pattern (5.1.4), but applied to the design
process instead. In a sense, design information updates our conception of the structure
design while still on the drawing board. The concept of using metadata to classify information about structure state is applied to the design process where the patterns can form
a metalanguage description that can help describe the design intent behind the structural
design information.

1.3.2

Chapter descriptions

A brief overview of each chapter is provided here to help the reader navigate the different sections of the thesis. Readers who want a top-level overview and some examples for
design might want to focus on Chapters 2 and 6. Understanding the vocabulary developed
in Chapter 3 will help the reader grasp the discussion of patterns and examples in Chapters 4 and 5, and the proposed extensions to performance-based design in Chapter 6. Each
chapter contains a final section giving a summary of the chapter's main accomplishments.
Chapter 2 provides an overview on how to incorporate the notion of performance into

10
planning for each phase of a structure's life from initial concept through its operational life.
This sets the context for applying the concepts discussed in the subsequent chapters.
Chapter 3 describes different types of structural design information and defines terms to
discuss them. The information content of a design changes with time as designers refined
the design by adding more information at greater levels of accuracy. Designs also mature
over time, replacing preliminary information that aids decisions with final information after
decisions are made. This chapter is presented first in order to provide the reader with
fundamental concepts about design that will provoke thinking about the design process
itself, instead of thinking about just one particular design or design method.
Chapter 4 describes ways to communicate the different kinds of design information defined in Chapter 3. The performance of the design phase can be thought of as the clarity of
the finished design. Poor designs are either lacking in information, or lacking in clear presentation. This chapter explains principles for clear design communication and discusses
the nature of communication. This chapter is by no means exhaustive, and provides just
enough breadth in order to explain the points in this thesis. Future research would expand
this chapter substantially.
Chapter 5 describes patterns of decision making in structural engineering and the information used to make those decisions. Again, this chapter is by no means exhaustive, and
intends to provide only the theory behind my contributions to performance-based design, in
a generalized manner. Future work on engineering design methodology could easily extend
the contents of this chapter.

11
The end of Chapter 5 walks through a sampling of common structural engineering design procedures and shows how the general patterns apply. The examples are intended to
provoke thinking about the design process itself and the thinking and judgment implied
by the calculation steps. The design methods may be presented in some textbooks as very
mechanistic, but they are in fact full of decision points where the engineer must practice
"best judgment" (Kardon, 2003). Seeing how the patterns apply in the common design procedures taught in books and articles will motivate the design process innovations explained
in Chapter 6.
Chapter 6 explains my extensions to performance-based structural engineering design.
The first example walks through the computation of post-earthquake repair time and repair
cost for reinforced concrete bridge design. The second example walks through an example
of incorporating construction performance into the design of a reinforced concrete structural element. The procedures employed makes use of the theory and concepts explained
in the preceding chapters.
Chapter 7 provides some closing thoughts about how the ideas in this thesis might
be applied to workaday structural design. I also indicate some directions for follow-up
research related to structural engineering design.

1.3.3

Patterns

Understanding the use of patterns to document and describe methods of design communication, analysis, and decision making will help the reader follow the arguments in this

12
thesis. Patterns are the essence of common solutions to common problems. In describing
the design process in this way, I hope that the reader will be able to pick up on my observations about design and understand where my proposed extensions to performance-based
design fit in.
Patterns describe procedures and practices for solving recurring problems in a general
way, which allows comparison and understanding of the tradeoffs involved in their use.
This "pattern language" or "design pattern" method was coined in the field of architecture
by Alexander et al. (1977), and has been applied elsewhere in the field of computer science
in describing object oriented programming techniques (Gamma et al., 1995). Chapters 4
and 5 use the pattern method to describe the processes inside structural engineering design
at a certain level of specificity needed to focus on the thesis examples, yet also with a level
of generality so that the same concepts can be applied in other domains.
Anwar et al. (2005) discusses the use of patterns in developing software for structural
engineering design. They develop the idea of reusable components for making structural
design decisions, but limit the discussion to implementation of of computer software. Similarly, Yu and Kumar (2001) describe the use of design patterns, but only for the specific
domain of finite element analysis. Also, Heng and Mackie (2008) describe five basic patterns for finite element analysis with an emphasis on object-oriented programming design
for analysis software. In this thesis, the method of patterns for description is applied to
the general design process itself, without a specific view on how to implement in software.
There is often a tendency to reduce problems down to computable components and only

13
work on those (Wing, 2006). But, it must be remembered that computers cannot think
outside the box, because they are a box. Computers cannot come up with something innovative because they are constrained by their own limitations. Rather, the human cognitive
process of learning and creativity is what must be pursued. For this reason, the patterns
in this thesis include the conceptual processes of communication and decision making and
not the computational processes of software and analysis.
The template used in this thesis to describe the patterns is based on Gamma et al. (1995,
p. 6-7) with modifications to fit the context of structural engineering design (Table 1.1).
Only the headings necessary for describing elements of structural engineering design information are used, and the descriptions for each heading are modified accordingly. Using
this common template makes comparisons between the different patterns more obvious.
The template also provides a method for recording the various trade-offs involved in using
each pattern in the design process.

14

Table 1.1: Pattern template.

Template section

Description

Pattern name

The pattern's name communicates the pattern's essence succinctly. The name is important because it becomes part of design
vocabulary.

Intent

The pattern's intent describes what the pattern does, what its rationale and intent are, and the design issues or problems it addresses.

Also known as

These are other common names for the pattern, if any.

Assumed

Description of the assumed background training of the reader

background

needed to correctly understand and apply this pattern. The description uses the categories of civil engineering knowledge defined in American Society of Civil Engineers (2008) at the bachelor's level, with additional references as needed.

Motivation

A scenario that illustrates the design issue or problem and how the
pattern is used to solve the problem. The scenario gives a concrete
example to aid understanding of the pattern's abstract description.

Applicability

The pattern's applicability describes the situations where the pattern can be applied, poor alternatives that this pattern addresses,
and guidance on how to recognize when the pattern is applicable.

15

Table 1.1: (continued)

Template section

Description

Structure

A diagram, illustration, or short example showing the a general


use of this pattern.

Consequences and

The pattern's consequences describe how the pattern fulfills its ob-

implementation

jectives, explains the tradeoffs and results of using the pattern, and
indicates what the pattern allows one to accomplish. Describes
any trade-offs, pitfalls, tips, or other information that designers
should be aware of when implementing this pattern. Also discusses any tool-specific information important for pattern use.

Known uses

Examples of this pattern found in real structural engineering examples.

16

Chapter 2
Performance-centric design
The contributions of this thesis are best applied in the context of achieving a performance-based structural design methodology that fully involves "performance-centric"
thinking throughout all the phases of a project. The performance of a structure during an
earthquake, for example, depends on its structural design. But the structural design does
not emerge from a vacuum, but rather as a product of the concept phase of a project. The
structure itself is a product of the construction and operation phases. Therefore, performance needs to be considered throughout the entire project delivery cycle (Figure 2.1).
This consideration of structural performance in all project phases is "performance-centric"
design. This chapter walks through the various phases of a project and points at places
where current research on design methodology can contribute.

17

Concept

Performance targets
and project scope

Design

Design information

Knowledge reuse
Construction

Operation

Repair

Figure 2.1: Project phases.

Building process

As-built state

Post-disaster state

18

2.1 Concept phase


The concept phase is where the general function, purpose, and use of a structure are determined. This is where structure performance must be first considered because this phase
establishes the general direction for a project. The general building topology, materials,
and general concepts for the basic structural system are determined here. Non-engineering
factors such as financing, architectural features, and permitting also happen here and if not
properly considered, could derail a project from efficiently fulfilling its performance goals.
To address these issues, research has been conducted with industry participants through the
UC Berkeley Project Production Systems Laboratory (P2SL). The documentation of successful case studies of structures (mostly buildings) has provided insight into the factors
that make projects successful.
The concept phase is the appropriate time for choosing the target performance of a
structure beyond service-level loads. This includes performance during earthquakes or
other hazard events. Determining the target performance requires adequate pre-event planning in order to incorporate the constraints on the designs and construction processes, and
indicate the appropriate needs for the operation phase. The structure owner should define
the acceptable risk of loss and risk of hazard exposure in service and extreme hazard conditions. For example, in nuclear structures, the acceptable probability of a core accident in
service conditions is set descriptively by federal law (10 C.F.R. 100). Risk management
consulting would be the most appropriate during this phase.

19

2.2 Design phase


The engineering design phase includes completing design information at the level of
detail needed for construction, and obtaining the permits necessary to begin construction
work. Three challenges to implementing fully performance-based thinking within the design phase exist today:
1. Producing analysis results that are meaningful to owners and other project participants is a difficult task that requires more supporting data and analytical procedures
than required for presenting the design to engineers, experienced builders, or a specialized audience who has expertise.
2. A disconnect often exists between what is designed in the office and what is optimal
for construction in the field.
3. An entrenched, point-based design process makes it difficult to explore design options for system-optimal structural solutions.
The first issue is addressed through the outcomes of the performance-based earthquake
engineering (PBEE) methodology developed through the Pacific Earthquake Engineering
Research (PEER) Center. This method of performance-based design enables consideration
of non-structural-engineering decision variables (DVs) by (1) extending quantification to
such decision variables, and by (2) providing a probabilistic acceptance criteria to determine whether performance is acceptable with an adequate reliability. The DVs could be
expressed in terms of dollar cost, material quantities, and probability. This analysis is per-

20

formed using analytical structure models subjected to earthquake ground motions, which
produce a relationship between hazard and demand. Then, that demand is correlated to
damage through empirical data. And finally, the damage is linked to repair methods, quantities, and costs based on historical and projected construction data. An example of this
method was developed for determining the repair cost and repair quantities of a California
reinforced concrete highway overpass bridge after an earthquake.
The second issue touches upon limitations in the process capabilities of contractors
and the interpretation of structural design communication, such as drawings, which are the
principal means of expression in practice today. An example of limited process capabilities
can be seen when designing sections with large amounts of congestion. Even if the design
seems to work on paper, installation in the field might be unfeasible due to tolerance issues and ambiguous details expressed in a 2D paper space. Or, the as-designed structure
might cost much more than expected due to hidden complexity. These issues are addressed
through design methodology research using the example of identifying factors contributing
to the notion of "congestion" in reinforced concrete structural elements.
Set-based design is proposed as a solution for the third issue, specifically for reinforced
concrete structures. Set-based design can be thought of as a funneling or narrowing process. An example of reinforced concrete building structures are provided in this thesis to
illustrate this general design method. The point-based design method goes from one trial
solution to another attempting to arrive at a final, feasible solution. The set-based design
method begins with the whole space of possible design solutions, and then intelligently

21
narrows the design set by making decisions at the "last responsible moment" (Ballard,
2000). Set-based design reduces rework and provides an opportunity to document why certain decisions are made. This documentation step helps to express original design intent,
and could be useful when designs need to change due to unexpected circumstances in the
field. When design intent is described in addition to a design solution, richer collaboration
can occur between engineers, general contractors, and rebar fabricators. Expressing design
intent demonstrates the implementation of decisions made in the concept phase.
All of the communication, design, and decision patterns have strong bearing during the
design phase. The communication patterns are especially important to implement properly because of the different levels of conceptual and concrete information communicated
during the preliminary and final stages of design. The analysis patterns show the different
types of analysis that need to take place during design. The patterns for decision making reveal different methods at choosing among various design options. The use of the
best decision patterns within design procedures can provide a solution with much greater
performance. Many design procedures focus solely on Engineering Performance (5.1.5)
without considering Social Performance (5.1.6).
Furthermore, the design itself can be measured for performance. The performance of
the design phase can be thought of as the clarity of the finished design. Poor designs are
either lacking in information, or lacking in clear presentation. The discussion of hermeneutics and patterns for communicating design information at the preliminary and detailed
levels explain methods for presenting designs in a manner appropriate the information.

22

2.3 Construction phase


Structural performance is not only an issue for the completed, final product, but rather
should also be a consideration throughout the construction phase of the project. During
construction, some demands are much higher than during normal service. For example,
mobile cranes and concrete trucks can impose significant demands on supporting columns
and foundations. Moreover, high quality construction is necessary to ensure that critical
details are properly built in the final structure. Therefore, it is of vital importance that a
full discussion of structural performance considers not only service and extreme events, but
also the construction phase of a project.
Describing design intent fosters a dialog between designers and builders which can
promote selection of more efficient details. For example, certain size bars specified in a
design might be more costly or difficult to install than other options. Cases studies and
interviews with structural engineers, general contractors, and rebar fabricators have helped
document these factors for structural design (Parrish et al., 2008a,b). Work flow simulation
has been performed using a Java web services implementation to experiment with the type
of information necessary to collaborate in a set-based manner (Wong et al., 2007).
Communication patterns requiring high maturity are the most useful here. For example, rebar placing information is essential for fabricating, delivering, and placing rebar for
reinforced concrete structural elements. A well-communicated design will aid construction
inspections by reducing the ambiguity in interpreting design intent among inspectors and
builders.

23

2.4 Operation phase


The operation phase of a structure allows verification of the structure's actual performance against the original design estimates. Previous work on sensors and state data frameworks provides the mechanism for capturing and querying performance data over time.
Using those frameworks, monitoring data can be made readily available for further use in
performance evaluation for the as-built state and post-hazard state, and estimation of future
performance related to long-term deterioration and degradation and estimates changes in
hazard exposure or service usage patterns. The major barriers to using the data involve
determining how to use the data for performance evaluation and identifying appropriate
questions to answer with the data.
One application of sensor data is to update a structure's demand model. A demand
model connects an engineering demand parameter (EDP) during an earthquake with a
given earthquake intensity measure (IM). Before an earthquake, the ground motion and
the structure's response are uncertain. After an earthquake, the IM can be determined from
seismograph records, and the EDP can be measured with structural monitoring sensors.
The observed response and expected response can be combined using the Bayesian update
rule to generate an updated demand model.
For a large earthquake, this method could be applied to post-earthquake rapid infrastructure assessment-like a "triage" for buildings and bridges. For moderate or small earthquakes, the data could be used to update performance estimates for future earthquakes.
The Updating pattern is useful here when applied to structural state data. The structural

24

state data from sensor monitoring information, or inspections can be used to update earlier
estimates on the structure's properties.

2.5 Repair phase


The repair phase of a project refers to the recovery process after a major catastrophe,
such as a large earthquake. The repair phase is typically not considered during initial
design for typical buildings and civil infrastructure, but current research is beginning to
address it. The LLRCAT methodology addresses the repair phase for reinforced concrete
bridges (Mackie et al., 2008c,a). Others have begun including the repair phase as part of
sustainability analysis during the design phase (Kang and Kren, 2006; Kneer and Maclise,
2008).
Special structures may have specific design procedures for repair and recovery after experiencing a hazard. Hospitals might be designed to need little repair in small earthquakes,
but allowed to require some repair in large earthquakes. Nuclear structures are designed
for safe shutdown and immediate re-start after a design-basis hazard, and to accommodate
emergency shutdown in an extreme hazard beyond the design basis.

2.6 Knowledge reuse


Lessons are inevitably learned from every phase of every project. The operation phases provides data from as-built behavior; the construction phase can provide feedback to

25
design; the design phase can capture the reasons for choosing one option over others; and
the concept phase provides the rationale for selecting performance objectives. Once performance is observed for a particular structure the data can be reused to improve future
designs. Future designs are often influenced through design code revisions which typically
require 5-10 years. Given this long feedback cycle, retaining design information is particular challenging. This learning process can be facilitated by using set-based design and
specifically documenting design intent.
In the case of retrofit, retaining knowledge of the original design intent and construction
documents would help engineers assess the structure's capabilities and options for retrofit.
The retrofit design process starts many years after the original structure is built and can
reuse much of the information from the original design.
Cost functions based on rebar reinforcement ratios and "congestion" in a joint is one
example of knowledge reuse. Data from the construction phase and performance data after an earthquake can be synthesized to enable efficient set exploration early in the design
phase. Research conducted through industry workshops has revealed the connections between communication during the design phase and the ability to reuse knowledge in the
future.
Knowledge reuse is encapsulated in this thesis through the use of patterns to describe
the concepts and procedures. The use of patterns allows one to reuse the methods and
understand why one certain procedure is more or less useful in different situations.

26

2.7

Summary

What I have done in this chapter:

1. Denned phases in the life of a structure by describing dominant activities in each


distinct phase.
2. Explained the importance of understanding performance in all phases.
3. Highlighted applications and research related to each phase.

27

Chapter 3
Information in structural design
This chapter defines terms for categories of information that express structural engineering design. These categories are used in the following chapters to explain the differences
between patterns of communication (Chapter 4) and the different emphases of the various
design analysis and decision patterns (Chapter 5). Using these terms helps clarify design
intent at the point of decision, and simplify discussion of how design variables converge to
their final values.
The discrete pieces of information are called variables because they develop with time
as the design process progresses. ATC-63 lists six basic design variables: occupancy and
use, elevation and plan configuration, building height, structural component type, seismic
design category, and gravity load. These design variables have related physical properties that express precise information (Figure 3.1). Additionally, Schon (1983, p. 98) describes "design domains", which encompass a larger range of design variables for building

28
Rt'laled Physic.il Properties

Typical framing layout

Distribution of seismic-force-resisting system


components

Gravity load intensity

Component overstrength

Distribution of seismic-force-resisting
components

Typical framing layout

Permitted vertical (strength and stiffness)


irregularities

Beam spans, number of framing bays, system


regularity

Wall length, aspect ratio, plan geometry, wall


coupling

Braced bay size, number of braced bays,


bracing configuration

Ratio of seismic mass to seismic-force-resisting


components

Occupancy and Use

Elevation and Plan


Configuration

Building Height

Structural Component
Type

Seismic Design
Category

Gravity Load

Ratio of tributary gravity load to seismic load

Story heights

Number of stories

Moment frame connection types

Bracing component types

Shear wall sheathing and fastener types

Isolator properties and types

Design ground motion intensity

Special design/detailing requirements

Application limits

Gravity load intensity

Typical framing layout

Ratio of tributary gravity load to seismic load

Component overstrength

Figure 3.1: Design variables for structure configuration have related physical properties
(ATC-63, p. 4-4).
projects including: program and use, siting, building elements, organization of space, form,
structure and technology, scale, cost, building character, precedent, representation, and explanation. These domains involve architectural style and artistic expression in addition to
structural configuration. Design information can be categorized by content using descriptors shown in Table 3.1 and by quality using descriptors shown in Table 3.2.

Table 3.1: Information content types answer different questions.


Content

Typical Questions Answered

abstract

how? parameters that are handy for design

functional

why? what are the reasons?

material

what is it made of?

process

who will make it? when will it be made?

spatial

where will it be?

Table 3.2: Information quality types answer different questions.


Quality

Typical Questions Answered

accuracy

does it hit the target?

ambiguity

how well are the information relationships known?

detail

what amount of precision is available?

grouping

what groups of pieces are referred to?

maturity

what will develop or change in the future?

precision

how close to the bull's eye?

30

3.1 Content of information


3.1.1

Material

Material variables describe an object's composition and intrinsic characteristics. This


includes whether the object is made of concrete, steel, wood, or some other material. It
includes the different grades of the material and any measurements on the composition and
engineering properties.

3.1.2

Functional

Functional variables describe the object's purpose and role in fulfilling the performance
goals of the structure. This information would describe that a shear wall serves to resist
lateral forces in a building, a beam carries the gravity load for a floor system, and that a
column carries gravity loads for a portion of the building. The functions of the individual
elements sharpen as the design process continues. Items exist in the design because of their
function in fulfilling the structure's purpose.

3.1.3

Spatial

Spatial variables describe an object's location, size, or shape. This includes information
such as the location and configuration of reinforcing bars inside a concrete structural element, the position of columns and beams inside a building, and the shape of a steel beam
or floor slab. The spatial information tells where the material goes.

31

3.1.4

Process

Process variables describe an object's method of construction or installation. This


would include phases and the sequence of construction along with any dependencies with
other pieces of the structure. This involves schedule information, procurement of the materials, positioning of the pieces, and guarantees that the built structure matches the designed
structure.

3.1.5

Abstract

Abstract variables deal with concepts that are handy for thinking and analysis. Abstract
variables are involved in the design process but are not necessarily needed for specifying
material and spatial properties nor for expressing function. For example, the reinforcement
ratio p is a design variable that is based on spatial and material information but is not
itself specifying the material composition or position, size, and shape of individual bars.
Other examples include the total cross-sectional area of reinforcing steel in a reinforced
concrete element, Aeff, the reinforcing index, co, and the ratio between repair cost and
initial construction cost, RCR. Abstract variables are typically derived from mechanics
models used for structural analysis.
Abstract variables can also be related to process information when dealing with measurements of construction performance such as PPC, the percent of planned items that were
completed.
"Performance goals" are abstract variables, because they are not intrinsically part of

32

a real object. Structural "performance" refers to a measurement of a structure's behavior


usually compared to an estimate based on a mechanics model. Performance could also
refer to the state of a structure during or after a specific catastrophic event (e.g. earthquake,
fire) or over a specified duration of time (e.g. 50 year performance). Performance then is
just another variable with different kinds of functional qualities. Structural performance
metrics apply at a specific level of detail, scale, abstraction, and aggregation.

3.2 Quality of information


3.2.1 Detail
Detail refers to the level of geometric specificity at a given scale. Low detail means that
very few of the spatial details are available. A 3D view or 2D drawing of something with
low detail would perhaps show the general shape of an object, but would leave out other
elements. Description with high detail can be viewed at a large scale and have features that
are visually prominent. A high level of detail is needed to make the designed object look
like what real object does. The highest level of detail is only achievable with the actual,
real object itself. Other representations of the object have less detail.

3.2.2

Ambiguity

Ambiguity is defined by Schrader et al. (1993) in two levels. Level 1 ambiguity is "Characteristic of a situation in which the problem solver considers the set of potentially relevant

33

variables as given. The relationships between the variables and the problem solving algorithm are perceived as in need of determination." Level 2 ambiguity is "Characteristic of a
situation in which the set of relevant variables as well as their functional relationship and
the problem-solving algorithm are seen as in need of determination." The concept of model
uncertainty in structural engineering (Ditlevsen, 1982) is related to level 2 ambiguity since
both relate to missing or neglected relevant variables.
These two levels of ambiguity function differently in a design situation. When someone is "figuring out a problem" it usually involves starting with a level 2 ambiguity and
working towards a level 1 ambiguity. Then, once that is determined, the solution method
can be chosen and the ambiguity can be resolved at the proper level of precision, detail,
and stability.
Most design methods work on the basis of level 1 ambiguities. They serve as a problem solving algorithm that provides reasonable values for each of the important variables.
However, one problem is that many design methods assume a point-based solution by only
allowing expression of variable values to a high degree of precision. Identifying the precision tolerances on these variables would leave set exploration more open.

3.2.3

Maturity

Maturity is the characteristic of design information that it has developed into a stable
form, close to its final value, and has enough precision needed for construction. Mature
variables have been informed by all the prerequisite information. In the literature, maturity

34

with respect to the amount of information content and variable values is referred to as
certainty or uncertainty, and maturity with respect to changes with time as stability. The
concept of maturity in this thesis refers to both aspects.
Schrader et al. (1993) defines uncertainty as "Characteristic of a situation in which the
problem solver considers the structure of the problem (including the set of relevant variables) as given, but is dissatisfied with his or her knowledge of the value of these variables."
Terwiesch et al. (2002) defines stability as "the likelihood of changing a piece of information later in the process." This also depends on the precision of the piece of information.
Low precision information is usually more stable, and high precision information is usually
less stable. When decisions are locked in as the design passes certain milestones, then the
locked down information is hopefully completely stable. A finished project is completely
stable, aside from information related to building operation.
Choosing a value that is highly precise at a stage of design that is too early might lead
to high uncertainty. Uncertainty involves not knowing the exact value of the variables at the
earlier stages of the design process. In some sense, progress during design involves labor
in order to reduce the uncertainty of the variables. Variables in the design are not always a
single numeric value. For example, the shape of an object might be a variable which needs
to be described in a more sophisticated way than a single linear dimension.
The term maturity is used instead "uncertainty" (Schrader et al., 1993) because it is
different in nature than that of statistical or model uncertainty (e.g. Ditlevsen and Madsen
(1996)). Using similar words would cause great confusion in a discussion of probabilistic,

35
performance-based structural engineering. Schrader et al. (1993) take the view that uncertainty is related to the design, or selection, of values that satisfy a certain set of criteria.
This is not the same as uncertainty regarding measurement, or uncertainty about a future
event, because in these types of uncertainty there is no design. It characterizes the gap
between limited knowledge and lack of detail. The concepts could be thought of a similar,
if one can imagine "measuring" a phantom object that does not yet exist spatially and materially. This is quite like what is done when describing a structure in a narrative. However,
the concepts are dissimilar enough to warrant different vocabulary for each category. New
terms are needed to distinguish these variables properly.
For similar reasons, maturity is preferred over "stability" (Terwiesch et al., 2002) because "stability" has different accepted meanings in the domain of structural engineering.
An stable structure is one that is resistant collapse, and an unstable structure is prone to
collapse. Neither sense of the word "stability" communicates the refinement of design
information and the development with time.
Furthermore, maturity is used only to describe design information, and not for the actual as-built structure and its operational performance. A mature design variable, such as
the specified length of a steel column, may actually represent a probability distribution
based on AISC tolerance specifications. The design variable is fully mature, having all the
information required for construction and expressing design intent, but the as-built condition may still be probabilistically uncertain to the designer. Aleatory uncertainty of the
structure in the sense of PBEE design methods considers the possible deviation between

36
the as-built structure and the mature design information used for construction.

3.2.4

Precision

Precision is defined by Terwiesch et al. (2002) as "the accuracy of the information


exchanged." As the design progresses the variables in the project will become more precise.
A finished project is completely precise because the real object contains all the spatial
information of the actual project. Numerical precision can be low with a range such as
10"-20" compared to a high precision range like 10.2"-10.4". Conceptual precision can
refer to generalities expressing purpose. For example the designer can use a lower precision
term like "lateral force resisting systems" and a higher precision description like "shear
wall". The combination of both numerical and conceptual precision is needed for project
completion.
Adequate control over the implied level of precision is important throughout the design.
If software tools are used to communicate design, there is often the mistake of allowing the
tool to imply a higher level of precision than intended. For example, it would make an
engineer nervous to see a rebar placer determining dimensions by using a scale and reading
off a number from a set of plans instead of using only the called out dimensions. This
is because the drawing itself is not intended to carry the degree of precision. It is the
dimensions which are supposed to express the precision. When a picture is drawn, it is
important to always carry along with it the text annotations that adequately describe the
intended level of precision.

37

3.2.5

Accuracy

Accuracy tells whether the variable's value is correct or not. It refers to how close a
variable's value is to its final state in the structure with reference to the implied precision.
In general, lower precision values are more likely to have high accuracy because there is a
wider range of possible values. Accuracy is important because it gives a measurement of
how likely values will change with time during the design process. This aspect of uncertainty in the final value is not the same as epistemic or aleatory uncertainty in the sense of
probabilistic hazard and structural performance, because it relates to the process of maturing design information and not the as-built condition of the real structure.

3.2.6

Grouping: agglomeration and aggregation

Agglomeration and aggregation are two different methods of grouping smaller physical
items into larger logical items. "Agglomeration" involves grouping dissimilar systems and
"aggregation" involves grouping similar systems (Yuan, 2008). For example, a bunch of
bananas is an aggregation of individual bananas, while a pizza is an agglomeration of crust,
cheese, and tomato sauce. Both the bunch and the pizza can be referred to as individual
items, but their underlying composition is different.
The rebar inside a reinforced concrete beam could be thought of as individual bars, or
as a single group of bars. This would be an aggregation of the individual bars into a group.
Similar objects are bundled together and considered as a single concept. The beam could
be thought of as a single object as well. The beam is an agglomeration of the concrete

38

portion and the rebar portion. Dissimilar objects are bundled together into a single object.
The issue of agglomeration and aggregation is distinct from that of precision. Precision
deals with how close the specification of the item is to its actual final outcome. Precision
operates at the appropriate level of grouping. For example, looking at a concrete beam
item to check clashes with mechanical equipment, specifying its overall dimensions to be
24" x 24" might be perfectly precise. But, if thinking about the individual rebar and concrete pieces as items, knowing the overall beam dimensions leaves a tremendous amount
of imprecision in the definition.
If the item being operated on is thought of at the group level, then it's constituent items
might be considered as a set of options because their individual specification is not given,
and might not yet be needed in the design process. For example, when performing structural
analysis on steel shear tab connection, the exact details of bolt spacing are not needed as
long as the agglomerate "connection" (consisting of nuts, bolts, and angle plates) can be
idealized as a non-moment-resisting pin. The exact bolts and plates to be used are left open
as a set of options until the group items need definition.
Another example is in the nailing configuration for a plywood shear panel. As long as
the nails can be treated as an aggregate group specified by a nail size and nominal spacing,
the exact position of each nail is not needed. Full spatial precision for design is obtained
for the "nail group" by giving the nominal spacing. But, full spatial precision is not given
for the individual nails, and is not required for construction. The exact position of each nail
can be left up to the installer. Good inspection evaluates individual nails for workmanship,

39
spacing, and size, at the same level of grouping as specified in the design.

3.3

Importance of content and quality

The importance of understanding the content and quality aspects of design information
are illustrated by looking at what is needed to define a mechanics model in order to inform
design decisions. Different qualities of information are needed to answer questions about
different parts of the structure. For example, bolt spacing might not be necessary to answer
initial questions about acceptable story heights, but will be necessary for the designer to
choose connection details, which may, in the end, impact the final story height through bolt
spacing and the number of bolts. The content and quality of information available during
the progression of a structure's design also influences the ways in which design information
can be communicated.

3.3.1 Mechanics models


Mechanics models are needed to determine whether a designed structure will perform
as expected before building it in reality. The mechanics models allows engineers to understand the behavior of something before it is built as a real object. The model allows
tinkering to determine the necessary design variables and relationships between them in order to resolve level 2 ambiguity. The model also serves as a mechanical abstraction of the
designed structure prototype. This gives the model the power to test design concepts on pa-

40

per (or computer) before the costly expense (or impossibility) of making the real prototype.
The model can quantify decision variables such that an acceptance criteria for performance
objectives can be applied. Most structures are too large and too complex to build samples
of the real-thing because off the one-off nature of construction projects in general. Before mechanics models, complex structures could only be made by master builders who
intuitively understood mechanics based on experience in constructing previous prototypes.
The mechanics models employed require a certain amount of information in order to
function. The different information content types are important for the mechanics models
because they must all be present in order to make a practical model. They generally require
the material information enough to specify the stress-strain behavior of essential components. The functional information must be known in order to determine what must be considered in the model and what may be ignored for the sake of simplicity. Enough spatial
information must be known in order to determine the location of the structure pieces, and
how the pieces interact with one another in compatible deformation. Process information
must be known in order to determine when different pieces are present in the structure during construction, and whether the structure can maintain equilibrium even if certain pieces
are not yet built. The abstract variables are the ones that are needed to run the mechanics
model and are commonly the results of the model that then need to be applied to inform
engineering design judgment in order to proceed in the structural design.
The quality of information is also important for mechanics models. The right level of
detail is necessary because if there is too little detail, then it will not be possible to define

41
a meaningful mechanics model of the structure. On the other hand, too much detail might
be unnecessary because some detail might not be included in the mechanics model at all.
This occurs when the abstract variables to be determined with the mechanics model are
not dependent on information at a too higher level of detail. The ambiguity of the design
information influences the ability to construct a mechanics model in the first place. In a
situation with level 2 ambiguity, the engineer might be unable to construct a relevant mechanics model because the problem solving method is unknown and therefore would not be
able to determine what information must be gleaned from the model. For these cases, similitude relationships may help by eliminating unnecessary variables, agglomerating groups
of similar variables (independent variables representing dimensions), and pointing to functional relations among independent variables through Il-terms (Sonin, 2004; Harris and
Sabnis, 1999).
The level of maturity of design information might also be affected by the mechanics
model depending on its complexity and the amount of information needed to compute and
construct it. Mechanics models usually operate at a specific level of precision, and when
that changes then the outputs will also be affected and will most likely influence changes in
the downstream decisions that depend on the model results. Immature designs are also less
likely to permit more detailed kinds of mechanics models. Immature designs might be wellsuited for very simplified types of models that capture the general global behavior of the
structure, but might miss out on some of the local effects that require more mature designs
and higher precision. Typically capturing local effects requires more complex models and

42

therefore require more of a time investment into creating and computing them. So, it might
be undesirable to commit to working on a mechanics model unless the design is more
mature and ready for detailed mechanics-related analysis.
The issues of grouping, substructuring, and similitude are relevant to mechanics models
because they determine what model elements are necessary. For example, in analyzing the
global behavior of a concrete moment frame building, it might be convenient to agglomerate the components that make up the columns (concrete, longitudinal rebar, and transverse
rebar) into an element that represents the column as a single prismatic frame element with
a stiffness EI and plastic moment capacity My. For a woodframe shear wall, it might be
good to aggregate the individual nails in the plywood panel so that it behaves as if the
panel were just rigidly connected. In connection design, it might be useful to agglomerate
all the different connection pieces and model them as a single semi-rigid connection object
idealized using springs.

3.3.2

Communicating quality of information

Clearly articulating the level of detail and accuracy necessary to define a design option
set at a given point in time during design requires open communication and understanding
of the values each party can bring and constraints that affect them. Lack of clarity on these
is an obstacle to set-based design. The right questions need to be asked at the right stage of
project design. Each project participant must understand not only what is asked, but also
the level of detail and accuracy that is required. Too much too early forces unrealistic and

43

undesirable commitment, while too little encourages rework.


The issue of level of detail appears in a variety of different contexts. In sensors and
health monitoring applications the information obtained from sensors can be found on the
local, element, and global levels. Estimating a global level parameter using only local level
data causes problems with data applicability. Estimating a local level parameter using only
global level information introduces too much uncertainty in processing the data. The right
information needs to be present in at the right level of accuracy in order to produce the
desired result. Different types of information at different scales have their own costs and
uncertainties.
In the design process, the issue manifests differently throughout the different stages.
In a sense, the stages of design themselves can be differentiated by the amount of detail
available for precise expression. At conceptual design, very few high accuracy details can
be expressed. By the construction phase, precise details are already clearly expressed in
the design.
There is a tension between freeform and parametric descriptions of structural systems,
elements, and details. A parametric description enables quick changes and easy use of
computer modeling tools. But, it also limits the ability to design something that is outside
the range of choices. For example, choosing to use an innovative structural system doesn't
easily fit in any conventional parametric model. A freeform description is necessary for
something new. Designing an innovation requires both creation of new objects and details,
and creation of new ontological categories to describe its function.

44

3.4

Summary

What I have done in this chapter:

1. Defined different aspects of the content and quality of structural design information.
2. Explained how different types of content and quality influence mechanics models.
3. Described challenges to communicating information quality.

45

Chapter 4
Design communication patterns
The clarity of communicating design intent is a performance goal for the concept and
design phases. The goal of communication is to transmit the intended level of information
without implying additional information and without neglecting any information. Good
communication instructs the receiver in the fullest detail with the greatest possible clarity
to achieve the purpose. When too little information is given, then recipient is left confused. When information is too ambiguous then erroneous assumptions are often made.
When communication is noisy with distractions then the information becomes obscured.
Confusing, cluttered, and noisy information presentation blocks successful interpretation.
This chapter describes the different patterns of technical communication for representing and transmitting design content. Using the most appropriate pattern of communication with at the correct level of information with the proper presentation design simplifies
understanding the details behind and intent of an engineering design. Each pattern has

46
different advantages in being able to articulate aspects of design variables themselves and
their interconnectedness to other pieces of information. The chapter begins with an introduction to the concept of genre as a communication pattern, and an overview of classical
hermeneutics. An understanding of hermeneutics allows analysis of the different patterns
of communication and the ultimate intent of conveying the author's meaning. After that,
the chapter discusses a list of various patterns describing their individual characteristics.
An example of communicating preliminary design information is given for the collection of estimates on bridge repair quantity, cost, and time (Mackie et al., 2008a). Specific
patterns of communication were necessary in order to communicate preliminary design information with bridge engineers at the California Department of Transportation (Caltrans).
Providing a mechanics model of the bridge was not helpful and neither were the spreadsheets containing the repair quantities. Rather, a different combination of communication
methods were best for conversation and building common understanding. The information needed was instead conveyed in damage scenarios containing (1) repair information
with spatial information, (2) dimensions and quantities, and (3) descriptions and names of
individual pieces.

4.1 Communication fundamentals and hermeneutics


Proper communication methods are needed to convey the information inside any building model. Different types of information need to be presented in different ways. The 3D

47
view of a building is not as good as a table for showing a list of materials to purchase, and
sometimes a 2D section cut can illustrate details better than a photo-real rendering with too
much unnecessary information. Each communication genre is characterized by a particular
form and purpose. Through the proper combination of appropriate methods, the engineer
will be able to communicate and document the design in the appropriate manner.
These different genres of communication are still necessary regardless of the manner
in which they are generated. Before the use of computers, the generation of drawings
involved laborious manual drafting. With the use of building information modeling (BIM),
the communication might be generated by extracting information from the building model.
Before investigating technical communication specifically, it is helpful to review existing, long-standing methods of communication in general. The various genres of communication have different strengths and are well-suited to different levels of abstraction. For
example, Fee and Stuart (2003, p. 22) list many different "genres", or patterns, of verbal
communication: narrative history, genealogies, chronicles, laws of all kinds, poetry of all
kinds, proverbs, prophetic oracles, riddles, drama, biographical sketches, parables, letters,
sermons, and apocalypses. Although all the genres of communication share common elements, the process of understanding each type can be drastically different because they use
a different form. Each genre also has different purposes. For example, a proverb or "rule
of thumb" has the purpose of communicating a general statement which may not be true
in all circumstances. A narrative history aims to explain the sequence of events and causes
leading up to a present condition. This genre might be most useful for communicating the

48
design process for a new hospital. The form and purpose of poetry might not be well-suited
for expressing construction details, but might be a suitable way of describing a building's
overall aesthetic impact on the community.
In a similar sense, there are different patterns of communicating design intent that are
appropriate for different phases in a project. The proper choice of a communication method
will assist the communication of the design and the design intent. Understanding these distinctions allows greater control over communication and allows the engineer to be more
clear, thus fulfilling the goal of helping the reader comprehend what originates in the mind
of the author. Different communication patterns are needed to adequately explain the concepts, spatial relationships, and information needed at different stages of design where
different decisions have and have not been made yet. Each genre is needed to express the
appropriate level of precision at the right stage in design. Several patterns of design communication are listed in Table 4.1 along with the information emphasis they are best suited
for.
A proper understanding of hermeneutics is important for communication, because of
the different concepts in design that must be clearly communicated. Hermeneutics is defined by Terry (1974, p. 17), who explains that:
Hermeneutics is the science of interpretation. The word is usually applied
to the explanation of written documents, and may therefore be more specifically defined as the science of interpreting an author's language. This science
assumes that there are divers modes of thought and ambiguities of expression
among men, and, accordingly, it aims to remove the supposable differences
between a writer and his readers, so that the meaning of the one may be truly
and accurately apprehended by the others.

49

Table 4.1: Communication patterns have different spatial, concrete, and abstract qualities.
Pattern

Information emphasis

Narrative

abstract, functional, material, process

2D Sketch

spatial

Annotation

associating of spatial and other types

Structural Details

spatial, material, functional

Spreadsheet

non-spatial

Mechanics Model

abstract

Rebar Placing

spatial, process, material for rebar

Schedule

process, with emphasis on time

A slightly broader definition is offered by Ramberg and Gjesdal (2005) who comment that
"hermeneutics covers both the first order art and the second order theory of understanding
and interpretation of linguistic and non-linguistic expressions." Non-linguistic expressions
such as drawings and figures are clearly also within the domain of hermeneutical application.
Hermeneutics becomes critical to the communication of abstractions in structural design because misinterpretation of design intent causes serious problems. It is in everyone's
best interest when all project participants aim to understand the original design intent. Hermeneutical principles exist in order to overcome the barriers to communication that exist
in the transfer of ideas. Understanding the principles identifies the areas where ambiguities

50
are common and highlights the issues that need to be overcome in the design process. An
understanding of the basic nature of communication helps the analysis of abstract communication in structural design. Thomas (2002, p. 27) lists the "time-honored definitions" for
hermeneutics, exegesis, meaning, and interpretation:
1. Hermeneutics is a set of principles.
2. Exegesis is an implementation of valid interpretive principles.
3. Meaning is the truth intention of the author.
4. Interpretation is an understanding of the truth intention of the author.
Challenges to the success of each process provide roadblocks to communication. For example, reading a drawing requires proper exegesis. From this perspective, one can understand
the general fear of novel computer tools providing an unintended implementation of exegetical principles, thus giving a wrong interpretation of the meaning. For example, many
engineers are somewhat wary of using electronic models as contract documents because of
the uncertainty in the ways information is expressed.
The goal in design communication is to accurately communicate the design intent all
the way through the project in a consistent manner. The handoff of information is where
comprehension of design intent can break down. Design intent is not subject to the reader's
interpretation. It is the writer's responsibility to communicate clearly, and it is the reader's
responsibility to apply correct principles in order to grasp the design intent properly. The
meaning of the design documents has its origin in the mind of the designer, and the goal
of handing off is to transmit that successfully to the next participant. Noisy communica-

51
tion can block that transmission. Doumont (2002a) notes three sources of transmission
"noise" in written communication: text (inconsistent paragraphs, complicated sentences,
faulty spelling), page layout (unclear structure, inconsistent typography, unusual fonts),
and illustrations (too many details on schematics, too many rules in a table, too many tick
marks on a graph). A good communicator would reduce the noise in all these areas.
A breakdown in communication results in problems. For example, interpretation might
be strained because of unfamiliar technical terms. Structural engineers are generally comfortable with discussing abstract quantities such as the reinforcement ratio p, which is commonly used throughout the design process. It is an abstract value based on a typical section
cut through a reinforced concrete element. Its value is not something that is calculated or
measured in reality. The code contains minimum and maximum values for p. Many design
equations use p and that number is used to help understand what is important to specify in
the design process. However, even though this quantity is important and well-understood
by designers, the quantity is largely unknown to rebar fabricators. This fact might give
the engineer pause to consider how to convey his meaning in a way that is more easily
interpreted.
Furthermore, Turk (2001) discusses the difficulties of conceptual product modeling
from a phenomenological perspective and his understanding of the process of the hermeneutical task. Product models are not themselves part of objective reality, but are created
from how the modeler chooses to describe objective reality. Therefore, Turk suggests that
there can never be a single unified product model that can capture all design information,

52
but rather that "several correct or different models may and should exist" and that new software "should not be built on a unified, centralized model, but on a combination of models".
Winograd and Flores (1986, p. 163) also note that "The most important design is ontologicaF, which is concerned with creating new types of things, not just creating instances of
similar objects. All the more, this points to the importance of carefully considering various
forms of communication to convey design intent. This indicates that any information tools
and BIM software for design must provide designers with the ability to control communication at the appropriate level of conceptual abstraction, and that no automatic system will
exist that take the place of human communication.
This issue is also seen in other domains, such as cartography. A well-made map will
provide information beyond mere spatial relationship and a mere photograph. In the same
way, the communication of design intent must be designed in a manner that allows for the
right manner of thinking, both symbolically and physically, by properly understanding the
different types and qualities of information. Imhof and Steward (1982, p. 79-80) comment
that:
The map is not only a picture; it is primarily a means of providing information. It has other responsibilities than has a painting. It must show not only
those features which appear in good light, but must also allow all similar things
to appear with the same emphasis wherever they are present. Above all, it must
include the depth of the valleys and the heights of the mountains, not only in
the visual sense but in the geometrical sense also. To do this one must transform some of these direct image effects and extend them through symbolically
indirect elements.
To this end, the following patterns of communication explain which ones are best suited
for different types of information.

53

Four general-purpose patterns are described first, then four more patterns that are more
application-specific. The patterns for Narrative, Sketch, Annotation, and Spreadsheet are
best for communicating freeform information. The Mechanics Model pattern is suited for
specific information related to structural behavior. The Rebar Placing Drawings pattern are
intended for the information needed by rebar placers and fabricators during construction.
The Database pattern is useful for standardized information. The Schedule pattern deals
specifically with time.

4.2 Patterns for communication


4.2.1 Narrative pattern
Intent
Describe preliminary conceptual information using word descriptions instead of using
sketches and drawings. Using words alone focuses on communicating functional, material,
and abstract information without prematurely communicating spatial information.

Also known as
Preliminary description, conceptual description.

Assumed background
Foundational + Design + Problem recognition and solving + Communication.

54
Motivation
During conceptual design, different options need to be explored and described without
too much detail. In the beginning of a project, the general building topology, material, construction sequence, and structural system need to be defined in a way that allows checking
for general feasibility without making premature commitments to details. This stage of immature design requires the ability to communicate functional information along with low
precision spatial and material information, and some abstract information that can be used
to perform basic structural analysis. Here, narratives are useful because they emphasize
exactly the types of information present in the conceptual design phase.

Applicability
Narratives are used in situations where some information needs to be presented in a
more conceptual manner, often at a lower precision than necessary for construction. A
word description is good at communicating design function without implying any premature spatial information. Material information can be communicated without making commitments to spatial decisions like shape and size. In the earlier stages of design, narratives
are preferred over detailed models with much 3D spatial information because of the ability
to describe preliminary designs at the appropriate level of detail.
A narrative should not be used when lots of spatial information needs to be communicated. Spatial data is difficult to extract from a word description, while a picture allows
the location and size of images to convey spatial relationships in a rapid fashion. For this

55
Line 2 and 3 at the Main Level (Plan 2/S3 on the structural drawings, the west and east walls of the stairwell
and hallway #1): The current scope of the remodel includes replacing the lateral load-resisting capacities of the
existing walls by adding plywood to remaining walls to replace plaster lost at the new openings. To upgrade the
remaining walls to the current code levels would require adding plywood at the alcove at the stairwell (by
eliminating the alcove, or by adding plywood from the living room side), and increasing the number of nails and
adding floor ties in the remaining walls.
Line 2 and!? at the Lower Level (Plan 1/S3 on the structural drawings, the west and east walls of the stairwell):
Upgrading this line would involve increasing the number of nails and adding other hardware at those walls
already set to receive plywood.

Figure 4.1: Narratives describe structural design at a functional level.


reason, narratives are often accompanied by a Sketch (4.2.2) because the sketch allows the
conveyance of spatial data without the precision needed for a model with 3D geometry.

Structure
Narratives often take the form of a short written report, or a letter. Narratives are often
accompanied by sketches that illustrate the spatial information for the items described in
the narrative. Figure 4.1 provides an example of a narrative describing the general scope of
work for a residential seismic strengthening project. The narrative explains basic functional
and material information without implying any more precision than is necessary at this
stage in the project.

Consequences and implementation


The narrative is able to fulfill its purpose of communicating information at a very specific level of precision because of the specificity and generality of words. Words can be
used to express a very specific detail, or can be used to express abstract categories and

56
types. An author can choose to use words to express the intent exactly. The communication can be done at the exact level of intended detail. The ability of verbal communication
to accomplish this strict control over level of implied precision is unparalleled in comparison with graphical communication alone. Doumont (2002b) comments that although
graphics are "indeed superior for conveying intuitive or global information. By contrast,
they do a poor job of expressing abstract concepts and lack the accuracy that words are
endowed with. Words, in a sense, are worth a thousand pictures."
Narratives are not suitable for communicating large amounts of complex spatial information alone. It takes many more words to express the same spatial information that can
be communicated with a graphic.
Narratives are able to explain the scope of a project with clarity that is beyond the capability of a sketches alone. Narratives typically may include a background or that discusses
project history, and the main purposes of a building project. Understanding the context
of the building helps to communicate the underlying intent of the design decisions and
options.
Because a narrative is generally brief, then the fact that something gets mentioned implies a significant weight. Each word in every sentence is intentional and chosen specifically. This is not the same as a 3D model inside a computer system where everything shown
on the screen is not necessarily intentional. For example, when a designer adds a steel beam
object to the model, there is usually no conscious decision to govern its appearance in every
conceivable view.

57
Known uses
Narratives are used when presenting a client with options for scope of work and general
design approaches. This allows the structural engineer to describe the nature and extent of
proposed work without committing time and effort to developing a full solution. The use
of a narrative helps build vocabulary between the engineer and owner, and can facilitate
communication at the functional level without needing to delve into spatial and material
design information. The narrative can also be useful in describing situations where the
construction process information controls the scope of work, such as in an existing building
remodeling project.
Narratives help describe general parameters constraining structural design. One case
study example from a medical office building contained the following sections:

1. Project description
2. Applicable standards and references
3. Material specifications
4. Geotechnical recommendations
5. Medical office building structural system description
6. Parking structure structural system description
7. Future bridge
8. Preliminary design sketches

The project description section includes the project location described by giving the

58

city, surrounding streets, and adjacent buildings. The building use is described as a medical
office building (MOB) and parking garage. The number of above grade and below grade
stories is given for the MOB and parking garage. Their orientation with respect to the side
of the project site is given along with the streets that the parking garage faces. Unusual
site conditions are described; there is a a natural creek that runs through the middle of the
site which requires special consideration. A statement of what is not included in the design
clarifies the scope of the structural engineer's role on the project.
The applicable standards section lists the standards used and the parameters assumed
for the schematic design. California Building Code (CBC) 2001 is used with parameters
for seismic design (zone, importance factor, source type, soil profile, near source effect)
and wind design (exposure class, basic wind speed, combined exposure factor for gust and
height Ce, importance factor).
The material specifications reference the applicable documents and specify certain parameters for use in the schematic design. The geotechnical recommendations provided a
summary of only the essential information required for completing structural design. The
structural system descriptions contain approximate dimensions for floor plan, roof penthouse, and basement wall thicknesses; general description of the construction work required to build the parking structure; approximate sizes of seismic joints between the main
building and parking structure; depths and approximate spacing of steel members for floor
system; and two different options for seismic force resisting systems. The future bridge
describes plans for a possible future addition. The design sketches display the spatial infor-

59
mation related to the narrative description. Figure 4.3 shows a floor system from this case
study.
Narratives can also describe acceptable hazard, risk, and performance objectives in a
manner that describes outcomes at the right level of precision. For example, the regulations
for nuclear reactor sites (10 C.F.R. 100) explain hazard and risk criteria for reactor site
selection, and also the desired performance objectives related to radiation exposure and
population risk. The regulations are written for general application to a wide range of
possible nuclear reactor types.

4.2.2 Sketch pattern


Intent
Describe preliminary conceptual information using sketches in order to show spatial information and relationships between objects without implying an artificially higher degree
of precision.

Assumed background
Foundational + Design + Problem recognition and solving + Communication.

Motivation
Sketches are useful for communicating spatial and non-spatial information together,
and are especially useful during stages of design when only preliminary spatial informa-

60
tion is available. The style of a sketch and the presence (or lack) of Annotations (4.2.3)
controls the amount of precision that is implied in the sketch. Sketches do not contain the
level of detail necessary to purchase and construct an item, but do provide enough conceptual information in order to understand function, purpose, and some degree of process
information.

Applicability
Sketches address the need to communicate preliminary design information in a way
that expresses essential spatial information, yet without implying more precision than is
actually needed during the current stage of design. Sketches are applicable even before
sufficient design information is available to use modeling software. The use of computer
models too prematurely during design might cause "fixation" (Jansson and Smith, 1991)
due to the structured requirements and overly precise geometry needed to use modeling
tools.

Structure
Sketches are typically used with Annotation (4.2.3), and often linked with a Narrative (4.2.1). Sketches are freeform graphics, and as such can be used to communicate
almost anything in a way that is best suited for the designer. Sketches can be custom tailored to suit any situation where conceptual information needs to be communicated, unlike
views generated from databases in BIM software which tend to be more rigid and difficult

61
to customize within the software itself. Figure 4.2 shows an example of a sketch pointing
out the location of a lateral force resisting system, without providing too much premature
detail. Figure 4.3 shows a sketch of a floor system inside a medical office building. The
sketch gives the spatial arrangement and depth of steel floor framing members and the deck,
but without implying a commitment to specific steel shapes.

Consequences and implementation


Lack of control over the level of detail poses a serious problem to the implementation of
BIM into the entire engineering delivery process. The views of a model are generated automatically by the computer software. Because of this, the engineer has less direct control
over the spatial relationships that are implied by the rendered views. Unlike a static CAD
drawing sheet, the dynamic view in a BIM system often is unable to contain consistent
annotation. In a static drawing sheet, the engineer can refine the graphic to express only
the significant spatial relationships in order to amplify their intended meaning. Furthermore, the static drawing can be annotated in the most appropriate way to express design
intent with accuracy at the appropriate level of detail. The engineer need to be concerned
about the drawing's view being changed thus corrupting the combination of text and precise
control over the exact spatial relationships shown.
In sketches, every stroke and annotation is intentional to the extent that they express a
certain thing in the designer's mind. Because of this, the conceptual specificity of sketches
can be more than a computer model view. A computer model view is a rendering of precise

62

0
i

-t

S>

Figure 4.2: Sketch shows general spatial information along with functional information for
preliminary structural system at an appropriate level of precision.

63

2-4"

lH.

3^"

^M/-C.

3" /ffz. p/^*-

Figure 4.3: Sketch shows the general depth and dimensions of floor system elements, but
without implying a commitment to specific steel shapes.

64
information already entered into some database system, while the conceptual sketch often
conveys information with a degree of imprecision in order to avoid premature specificity.

Known uses
Sketches are often used in communicating preliminary design information and can be
readily used even when the project is immature and highly precise design information is
not yet available.

4.2.3 Annotation pattern


Intent
Display information in a visual format consisting of a main graphic element with text
that explains the important components of the graphic.

Also known as
Call outs, notes, markup, labels, tags.

Assumed background
Foundational + Design + Problem recognition and solving + Communication.

65
Motivation
Graphic elements, on their own, lack the specificity needed to communicate what the
main point of the graphic is. Annotations link spatial and non-spatial design information,
and show which portions of the graphic are meant to convey information and which ones
are not. Annotations show material design information when items are called out, and can
explain functional information that would otherwise not be presentable in a graphic form.

Applicability
Annotations can be added to any graphic element in order to highlight and explain its
significant pieces of information. This addresses the increasing tendency to allow software
programs to generate drawings and models without seriously thinking about how to teach
the reader what information is significant. Annotations provide information at the point
of need and make communication more clear. When more information is provided, the
graphical communication becomes more clear and easily understood and used.
Annotations are useful when grouping needs to be communicated, or functional information that describes the intent of an item that goes beyond its spatial arrangement. The
text annotations can describe categories and define terms, and express other information
that is not easily represented in graphical form.

66

Figure 4.4: Annotations enable proper interpretation.


Structure
An example of the importance of annotation can be seen in three drawings in Figure 4.4.
Each drawing has the exact same graphical, but with different verbal annotations. The first
drawing has no annotation, the second shows geometry plus a description of the bar size
and bar count, the third gives the individual position of the bars. No single drawing is
"correct", but each conveys a very different information content. The center drawing is
perhaps the most clear in communicating the original design intent for the rebar placement
and selection. The additional information given in the right drawing is not necessary to be
specified, and might even be incorrect because it implies a level of precision higher than
necessary.
This example illustrates that good communication is sometimes necessarily redundant
(Doumont, 2002a), and that there are multiple different ways to model the information

67
content of a single design expression. The position of the rebar could be modeled by
giving the individual position of each bar relative to the concrete member, as depicted by
the third drawing. Or, the position of the rebar could be modeled at a greater level of
aggregation as in the second drawing which represents more of a conceptual layout instead
of a physical layout. For visualization, it may be necessary to compute an exact position
of each individual bar, but in reality the specification of the rebar might only be described
on the level of detail of the annotation in the second drawing. This leads to problems of
over-precision that occur when too much information is inferred that was not intended.
Consider how an engineer would feel if he saw a subcontractor measuring a drawing
with a scale to determine the spacing between essential structural components? It would
be slightly disturbing to say the least, because the drawings were never intended to be measured. Rather, the engineer used the drawings to convey spatial relationships and intended
for the specifically-expressed dimensions to control the important, defining characteristics.
If the exact position of each bar were really necessary and critical to the design, then drawing (c) would be used. If not, then drawing (b) would have been used.
This notion of level of detail and level of abstraction and aggregation becomes more
important in preliminary design. In the earlier stages, it is more important to remain open
and slightly "fuzzy" on the details. In the later stages of design, exact locations and detail become more important because different kinds of performance questions are being
considered.

68
Consequences and implementation
Annotations accomplish the communication of non-spatial information that is linked
to spatial information presented in a graphic. Annotations need to be carefully thought
out so that the meaning can be communicated clearly. Text can be added to graphics in a
number of different ways, by using an arrow or some other pointer, or simply by positioning
text in the appropriate location. As free form tools, annotations are inherently able to
communicate the appropriate level of abstraction and concreteness.
Implementing good annotations may depend on the drawing tools that are used by the
designer. Adding annotations to at hand sketch is obviously simple, but adding annotations
to a view from BIM software might prove complicated.

Known uses
Annotations are used everywhere in order to add information on top of spatial information in a graphic. It is used to describe non-spatial information that must be linked to a
spatial location on a graphic. Poor drawings that cause frustration and miscommunication
often have inadequate annotation.

4.2.4 Spreadsheet pattern


Intent
Present information in a tabular format in order to provide easy access to data that is
linked or sorted in a particular way using text and numbers.

69
Also known as
Tables, Excel sheets.

Assumed background
Foundational + Design + Problem recognition and solving + Communication, and general knowledge of spreadsheet concepts and tools, such as Microsoft Excel.

Motivation
Spreadsheets are useful because they are general purpose tools for communicating all
different kinds of information and can be accessed in a computer program. The spreadsheet
is able to perform logic operations and basic computation.

Applicability
Use spreadsheets when
1. One needs a general purpose tool to convey text or numeric information.
2. Data entered by the designer needs to be transformed through computation. The
spreadsheet can automatically perform some computation through cell references
and formulas.
3. Reference data needs to be looked up by a person and by a computer program.
4. Linking the information to spatial location is not significant.

70
5. Different types of information need to be linked and presented together, as in a table
showing two variables across the rows and columns.
6. The information can be located by using multiple coordinates. For example, the row
and column of the data in a table identifies a piece of information.
7. Multiple tables need to be displayed together or in series.

Structure
A spreadsheet displays data in a tabular format. This permits a free-form logical arrangement that can be specifically suited for a work task. The advantage of being able to
show only specific pieces of information helps make communication clear. However, this is
also somewhat of a disadvantage since spreadsheet tables often lack the spatial linking that
is often necessary to understand the data in the context of the whole structure. Figure 4.5
shows an example of a spreadsheet containing the unit repair costs for bridge repair items.

Consequences and implementation


An advantage of using spreadsheets is the freeform ability to present and store information using a view that is the most needful for the specific application. The ability to
implement some logic within the spreadsheet is also a particular strength.
The barriers to implementation are extremely low because almost everyone knows how
to use and edit a spreadsheet and has access to spreadsheet software, such as Microsoft
Excel. The ability for most people to edit the information and customize it without having

71

A
1 Unit costs associated with each repair quantity
2

Item Name
Structure excavation
Structure backfill
Temporary support (superstructure)
Temporary support (abutment)
Structural concrete (bridge)
Structural concrete (footing)
Structural concrete (approach slab)
11 Aggregate base (approach slab)
12 Bar reinforcing steel (bridge)

Unit
CY
CY
SF
SF
CY
CY
CY
CY

13 Bar reinforcing steel (footing, retaining wall)


14 Epoxy inject cracks
15 Repair minor spalls
16 Column steel casing
17 Joint seal assembly
18 Elastomeric bearings
19 Drill and bond dowel
20 Furnish steel pipe pile
21 Drive steel pipe pile
22 Drive abutment pipe pile
23 Asphalt concrete
24 Mud jacking
25 Bridge removal (column)
26 Bridge removal (portion)
27 Approach slab removal
28 Clean deck for methacrylate
29 Furnish methacrylate
30 Treat bridge deck
31 Barrier rail
32 Re-center column

LB
LF
SF
LB
LF
EA
LF
LF
EA
EA
TON
CY
CY
CY
CY
SF
GAL
SF
LF
EA

3
4
5
6
7
8
9
10

LB

""f
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$

tiean
165
220
38
38
2,225
520
1,625

$
$
$
$
$
$
$

std dev
33
44
8
8
445
104
325

Notes
$250for<50CY
$335 for < 14 CY, $ 1 1 5 > 2 5 0 C Y
placeholder
$2000 for > 10 CY

325

1.35

65
0.27 $6.50 < 400 LB

1.20
215

$
$

0.24 placeholder
43

300
10.00
275
1,500
55
55
2,050
9,000
265
380
3,405
2,355
1,000
0.40
85
0.55
2
100

$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$
$

60
2
55
300
11
11
410
1,800
53
76
681
471
200
0.08
17
0.11
0.40
20

s1 has100/sf
MR101-160mm

Class 625 (140)


610mm diameter. Class 625.
610mm diameter. Class 70.

placeholder
placeholder

Figure 4.5: Unit repair cost spreadsheet (Mackie et al., 2008a).


specialized skill is a tremendous benefit for interaction.
The liability of using spreadsheets is the lack of spatial information relating the information to a position within the structure being designed. This poses a challenge to
understanding data in the proper context. For example, rebar bar list might be useful as a
spreadsheet for estimating purposes, but would be a poor way to show design information
for installation because of the lack of spatial connectivity. Using spreadsheets for recording
the repair quantities in bridge earthquake damage scenarios was not as good for communicating with a cost estimator as was placing the same information on drawings Mackie et al.
(2008a).

72
Known uses
The use of spreadsheets is ubiquitous in engineering calculations. They are useful in all
circumstances because of their free-format ability to communicate the right type of design
information with the proper accuracy.
Spreadsheets are used commonly used for documenting design procedures involving
many computations. Christy (2006) provides examples of how to use spreadsheets for performing and document structural engineering calculations. The examples cover basic mathematical operations such as integration and matrix math, and detailed design examples for
loading, connections, foundations, storage tanks, slabs, beams, and columns. Spreadsheets
can provide a "clear audit trail and uncomplicated code" and are useful in documenting calculations inside a single document. The ubiquity of spreadsheets allows them to function
as a universal platform for communicating information.
Spreadsheets are used to implement the database described in Mackie et al. (2008a)
for reinforced concrete bridge repair cost and time. A series of spreadsheets contain basic
information about the material and geometry of the bridge, its analyzed performance subjected to multiple earthquake ground motions, and data on repair cost and time estimates.
The spreadsheets are used to implement a database structure needed to store and access all
the information needed to support the methodology. In this example, the spreadsheets are
linked to each other using cell references so that changes in one spreadsheet update in the
other spreadsheets. It serves as both a data storage mechanism, and an easily programmable
data entry method.

73

4.2.5 Mechanics Model pattern


Intent
Use an abstract model of a whole structure, or structure components, based on mechanics theory to show the structure's response to loads and deformations.

Also known as
Analysis model.

Assumed background
Foundational + Design + Mechanics + Materials science + Problem recognition and
solving + Communication.

Motivation
During preliminary design, it is often necessary to explore the feasibility and communicate the essential nature of a structural system. Describing how a structure works using a
mechanics model communicates the internal characteristics of a building and reveals how
the structure will hold itself up and resists its loads. The mechanics model is able to show
a functional view of how system components interact, because the mechanics model is abstracted. This makes a mechanics model suitable for communicating preliminary functional
information along with basic spatial and material information at the same time.

74
Applicability
A mechanics model is useful for showing the essential features of a design that characterize its structural behavior and demonstrate how it will perform its function. This does
not necessarily refer to computational mechanics model which is computed using software,
or a method suitable for hand calculation. Rather, the idea in this pattern refers to the way
that this information is communicated in the design process. For example, even a simple
force flow diagram, or deformation mechanism could be an appropriate mechanics model
for the concept phase of design.
Using a mechanics model to communicate design intent avoids the unintended effect of
premature precision when the design is still immature. It avoids the problem of implying
additional details that have not yet been designed. The mechanics model also makes clear
which parts are relevant to structural performance, and which parts are not. This distinction can reveal areas where other design changes can be made without affecting structural
performance.

Structure
Mechanics models are generally shown using a schematic diagram combined with numerical data supporting structural performance and spatial information. Mechanics models
tend to show less detail than other forms of communication because they are highly abstract
in nature. An example of a simple 2D mechanics model with a nonlinear pushover curve,
moment demands, and displaced shape is shown in Figure 4.6.

75

Nonlinear Static Pushover for Frame 1

1200 h

4
5
6
Horizontal Roof Displacement

Figure 4.6: Mechanics models show essential design features and structural performance.

76
Consequences and implementation
The implementation of a mechanics model requires a simplification of the actual structure. Only the components that influence the outcome of the model are considered, and
other information that is not relevant to the abstract variables that come from the model are
generally ignored or not used. Much wisdom and practical experience is needed to judge
what aspects to include or exclude, and what level of accuracy and precision is sufficient.

Known uses
Analysis models are appropriate for showing abstract variables. The analysis model is
created for the purpose of manipulating objects related to structural engineering, mechanics, and materials theories. They are well-suited for communicating the non-real variables
that are needed to make engineering decisions that are based on mechanics and calculation.
Of course, not all decisions are based on analysis. In fact, the limitations of the analysis
model for decision making highlights places where mechanics theories are insufficient for
completely describing the factors necessary for making the design decision.

4.2.6 Rebar Placing Drawings pattern


Intent
Provide drawings with information necessary for a rebar placer to tie and place reinforcing bars within concrete formwork.

77
Also known as
Sometimes inappropriately known as "shop" drawings.

Assumed background
Foundational + Design + Problem recognition and solving + Communication + Teamwork, and either familiarity with the CRSI design and detailing manuals (Concrete Reinforcing Steel Institute, 2000,2001, 2002, 2008) or experience comparable to an apprenticelevel ironworker.

Motivation
When an ironworker is out in the field, he has little time to analyze a set plans or compile
information that is found in difference places. All of the relevant information needs to be
readily available down to the level of individual bar locations. The information must also
include all the process information and enough spatial information to identify and place
each bar. This includes details such as bar chair location and intermediate supports for
stability.
Good placing information will locate all the relevant information the point of need. This
involves using graphics and appropriate annotation to explain what each graphic element
indicates.

78
Applicability
Placing drawings are used to communicate a fully mature design that has sufficient
precision for construction. These drawings emphasize spatial, material, and process information for rebar, and abstract information is largely absent. These drawings show a tight
linkage between the spatial and material information by showing the location of each individual bar and indicating the bar size, type, and configuration of each one. This close
coupling of information makes the placing drawings particularly useful to field ironworkers.
Placing drawings are an example of a communication pattern that is inappropriate for
preliminary design. The required amount of precision is too high for an immature design.

Structure
Placing drawings are typically created using general purpose CAD software augmented
with special tools for drafting rebar. They typically make heavy use of cross section and
elevation views and use annotations to describe the individual bars. Intelligent grouping
of the bars makes the drawings easier to read. The level of detail is much higher than
that of other drawings because placing drawings need enough precision for construction
(Figure 4.7). They must enable measurement and provide fixed reference points that can
be identified in the field, which allow placers to determine the location of individual rebar
pieces.

79
- i - as i r-i" TOP
4- MS X S'-O* LEFT SHE
2 - 16118 giCKT S!K

Mowspaw

^ u JJfi?- 2SIB

"WEFT

StSECTM

.4-

SSMf

(b)
Figure 4.7: Placing drawings (a) show each individual bar, but other drawings (b) of the
same object might only show aggregations of bars (Concrete Reinforcing Steel Institute,
2000, p. 17-14,17-15).

80
Consequences and implementation
Placing information is typically communicated using a set of 2D drawings printed on
paper. The use of paper is good because of portability and the ability to see a large amount
of spatial information at once, providing context for the information presented.
General purpose drawing software is perhaps best suited for creating placing drawings
today. However, recent versions of Tekla Structures now provide a high enough level of
detail for modeling individual rebars. Currently, the placing drawings are created by detailers, trained technicians who specialize in making placing drawings based on structural
engineering design information. Their expertise includes the addition of information that
makes it easier for an ironworker to perform his job well.
A barrier to implementation is the fact that the level of precision needed for generating
placing information is much higher than that typically needed for structural design tasks.
Placing drawings are "an entirely different matter" from other drawings, because they require the shop drawings of other trades, exact dimensions for concrete formwork, and the
precise construction sequence (Birley, 2008). The creation of placing drawings also reveals
areas where the design might not have the level of maturity necessary for construction. This
is because the level of precision needed to develop a mechanics model, estimate costs, and
estimate quantities is much less than that required to make placing drawings. Information
such as the location of bar chairs might not be needed to design the structural components,
but are required to build them and must be shown on placing drawings.
This requirement for more information poses a challenge for applying performance-

81
based design concepts that extend to placing information. During the earlier stages of
design, there is not enough information to generate placing information. Yet, it would be
helpful to know how preliminary design decisions might influence the overall ability to
place rebar.

Known uses
Placing information is ubiquitous in engineered projects made of reinforced concrete.
In current practice, placing drawings are typically created by rebar detailers. Placing drawings are defined by several different industry bodies.
Concrete Reinforcing Steel Institute (2002) defines placing drawings and draws a clear
distinction between placing drawings and the inappropriate term "shop" drawings:
Placing drawings are working drawings akin to erection or assembly type
drawings, instructing the field Ironworker (Placer) where to place and tie the
reinforcing bars within the formwork. Placing drawings may also indicate
the bar support layout with a placing sequence, thus facilitating the efficient
installation of the reinforcing bars.
Placing drawings are prepared by detailers, trained technicians who are not
necessarily graduate engineers, but are extremely proficient in interpreting the
structural information shown by the contract documents. At no time does a
Detailer make an engineering decision. In fact, in today's litigious society, a
Detailer employed by a Reinforcing Bar Fabricator would be foolish to accept
this responsibility.
Detailers also prepare lists of the reinforcing bars that are shown on the
placing drawings. A bar list is a listing of reinforcing bars making up a bill
of materials. The bar list contains the quantities, sizes, lengths, and bending
dimensions of the reinforcing bars. These lists serve several purposes. The
Fabricator uses a bar list for shearing and bending, tagging, shipping, and invoicing. The Ironworker Foreman and the placing crew use a bar list for checking delivery quantities, sorting bundles of bars in the job-site lay-down area,
and hoisting of proper bundles to the placing area on the formwork.

82

Placing drawings are not used in the fabricating shop, per se. Thus, the
generic term "shop" when applied to the detail drawings for reinforcing bars
in site-cast concrete construction is extremely inappropriate.

ACI 315, Sec. 3.1-3.2 mentions specific items that should be in placing drawings:
Placing drawings are working drawings that show the number, size, length,
and location of the reinforcement necessary for the placement and fabrication of the material. Placing drawings may comprise plans, details, elevations,
schedules, material lists, and bending details. They may be prepared manually
or by computer.
Placing drawings are intended to convey the Engineer's intent as covered in
the Contract Documents... The placing drawings must include all information
necessary for complete fabrication and placing of all reinforcing steel and bar
supports.

American Society of Civil Engineers (2000, Sec. 17.3.7) defines placing drawings as
essential for safety and performance and discusses the design professional's responsibility
to approve placing drawings:
Concrete placing drawings illustrate the reinforcing steel components that
will be part of a completed structure. These components are crucial to the
safety and performance of a completed facility.
Reinforcing steel components are furnished and placed according to the design professional's specifications in the construction contract documents. Examples of components for which placing drawings are prepared include castin-place concrete and post-tensioned pre-stressed concrete structural elements.
The design professional has authority and responsibility for overall design
of the completed structure and for the review and approval of the placing drawings for conformance with the project design concept and the information in
the construction contract documents.
The constructor and subcontractors have responsibility for preparing the
placing drawings, providing the materials specified, and completing the fabrication and construction processes. This work is carried out in accordance with
the construction contract documents, approved placing drawings, and accepted
industry standards.
In most cases, placing drawings for reinforcing steel in cast-in-place concrete do not need design services and it is not necessary or appropriate for

83
the contract documents to call for certification by a registered professional engineer. For post-tensioned pre-stressed cast-in-place concrete structures, the
design professional may delegate certain design activities to a specialty engineer employed or retained by others and provide specifications for the loading condition and other design parameters in the contract documents. In such
cases, the design professional retains responsibility for the overall safety and
performance of the completed structure. The specialty engineer is responsible
only for the design work delegated and certifies with signature and seal that
the related calculations and drawings meet the specifications provided by the
contract documents.

4.2.7 Database pattern


Intent
Store structure information in a database so that it can be reused by a computer program
for further analysis.

Assumed background
Foundational + Design + Problem recognition and solving + Communication, comprehension of relational databases (e.g., Codd, 1990), and knowledge of the capabilities of
BIM software for structural engineering (Eastman et al., 2008).

Motivation
When information is stored in a database, it can be accessed programmatically by computer software. This is useful in BIM applications where communication methods such as
drawings, lists, and 3D renderings are constructed as views of database content. Databases

84

are also useful when tools for structural analysis are used to interact with other software
for design and coordination. Using a shared database avoids data duplication by requiring information to be entered in only one location and used in several other applications.
This also reduces errors because a single change can be made once for the entire database
instead of multiple locations.

Applicability
Databases are applicable when the information needs to be retrieved by another computer program, and when where the input data are already in a known form. This does not
work as well when information is of a free form nature, or is in a preliminary stage where
it cannot be defined with the precision necessary to fit into the database schema. In those
situations, it is better to use other communication patterns such as the Sketch (4.2.2) and
Narrative (4.2.1).

Structure
A database requires a data model that expresses the form of the information content to
be stored in the database. A relational type model is commonly used, representing data in
the form of tables and views (Codd, 1990). The relational data model for reinforced concrete bridge structural performance, earthquake hazard, repair methods, costs, and damage
states illustrates the Database pattern (Figure 4.8). This example shows the relationships
between the different components of the LLRCAT loss estimation methodology presented

85
in Mackie et al. (2008a).

Consequences and implementation


Using a database requires the definition of a data model to store all of the information.
Creating this data model could be very difficult when handling different types of information content, especially when the different data types are combined with one another
at different levels of grouping. The relationships between different pieces of data are not
always well-defined, particular with functional design information, and for designs that are
at a lower level of maturity.
Defining the views needed to communicate database content is an additional step of
work that would not be needed if the communication were created directly. This would,
of course, depend on the kind of information that is already stored within the database.
These views must also implement a method for entering information into the database,
which might require substantial programming overhead in order to store information into
the appropriate fields.
Data validation is another barrier to implementation. If the data is assumed to be in a
particular format or type to be used by a computer program, proper care must be taken to
ensure that all of the data is valid before entering the database.
Databases might be inappropriate in cases where the rigidity of the data model prevents
certain objects from being properly communicated. In these cases, a freeform communication method such as a Narrative or a Sketch would be more appropriate. Databases might

86

Bridge Information
PK

info id

Downtimes

name
value
unit
metricValue
metricUnit

PK

time id
downtime
downtimeSD

-has-

Damage States
Performance Groups
PK

PG id

FK1
FK2

location
DS_id
EDP_id

has-many-W PK
FK1

has

Repair Methods

PS..Jd

FK2

PK

repair id

FK1

qty_id
repairDescription

1..*
"has many

repairjd
name
lambda
beta
description
state
time id

V
Repair Amounts
PK

EDPs
PK

EDP id

FK1

unit
description
result_id

Performance Results
1

nany-^

PK

result id

FK2

INUd
EDPvalue
EDP_id
IM_value

FK1

Production Rates

Repair Quantities

PK

PK

QJd

FK1
FK2

itemName
rate_id
cost id

rate id

amount id

Intensity Measures
1

fe,PK

IM id

amount
repair id
Q_id

unit
description

Unit Costs
PK I cost id
-ha:

unit
unitRate
unitRateSD
notes
PERTmin
PERTmax
PERTmode

-has-

unit
unite ost
unitCostSD
notes

Figure 4.8: Relational data model for bridge performance database.

87
also require a certain amount of precision that make it unsuitable for communicating preliminary information when the designs are not mature enough to support such precision.
A single model is always unable to capture all relevant information, and the best use of a
database would be to combine it with other communication patterns to express the whole
design.

Known uses
Software for BIM commonly are examples of the Database pattern to communicate
general design information. Currently, the two most prevalent commercial programs for
structural engineering applications are are Autodesk Revit, and Tekla Structures. These
computer programs generate views of the information stored internally as a database, and
are intended to allow external programs to interface with the design information stored
internally.

4.2.8 Schedule pattern


Intent
Display process information with an emphasis on showing actions linked to time so that
concurrency and dependency can be easily seen.

Also known as
Timeline.

88
Assumed background
Foundational + Problem recognition and solving + Project management + Communication, and knowledge of the Lean Project Delivery System (Ballard, 2001).

Motivation
The schedule pattern is useful when the duration of a construction project needs to be
estimated. This allows communication of time dependencies and the sequence of activities.
Schedules constructed at different levels of detail allow the communication of different
information. During preliminary design phases, the schedule can show general feasibility
and milestones for design and construction.

Applicability
A schedule is one of the best methods for showing process information. It addresses the
problem of communicating how something will be built. The material and spatial information needed to describe the final state of the structure can be broken down into construction
steps and shown in a schedule. Schedules communicate changes in material and spatial information with time during construction. This pattern addresses the poor alternative of only
describing the final state of the product without due consideration of the process behind it.

89

ID
1
2
3
4

Task Name

Start

Finish

3/5/2008 3/5/2008
Excavate Bent 5
Set Column Casing and Weld 3/6/2008 3/7/2008
Grout and Paint
3/10/2008 3/17/2008
Backfill Bent 5
3/18/2008 3/18/2008

Duration
1d
2d
6d
1d

Mar 2008
5 | 6 | 7 | 8\ 9 \10\11\12\13\14\15\16\17\18

^am

Figure 4.9: Schedule for bridge repair scenario task (based on Mackie et al., 2008a, p. 177).
Structure
Schedules can take many different forms depending on depending on the method of
scheduling and the intent of the schedule. Schedules can be represented at different levels
of detail. A schedule at a lower level of precision might only show general tasks for the
overall project, while a schedule at a high level of precision might show individual tasks
for individual workers.
An example schedule based on an estimate for a bridge repair task is shown in Figure 4.9 (based on Mackie et al., 2008a, p. 177). This example provides information at the
level needed by Caltrans to perform estimates for budgeting.

Consequences and implementation


Implementing a schedule can be difficult because the process data needs to be expressed
as activities with time, durations, and sometimes also resources. The dependencies of
activities must be known and whether the activities will happen concurrently, or one at a
time. There might not be enough information to estimate a schedule unless the process
information for the design is mature enough.
Schedules usually have difficulty presenting information that is spatial in nature because

90
the focus is not on where, but on when. This deficiency can be overcome using particular
methods of showing schedules for repetitive work, such as the Line of Balance method
(United States Navy, 1962; Al Sarraj, 1990), or concentrating on site planning
Schedules need to be created at a particular level of grouping in order to maintain
clarity. The different units of work are broken down into appropriate levels depending
on how detailed the work plan needs to be in order to satisfy the user of the schedule.
For example, when doing initial preliminary planning to investigate the feasibility of a
structural system option, the exact details for when and where to install dry wall panels
might not be necessary to know.

Known uses
Schedules are used in all construction projects, and especially in the construction phases. Two common formats for showing scheduling information are the Gantt chart (Figure 4.9) and the Line of Balance (LOB) method. The LOB method was developed by the
U.S. Navy during the 1950s (United States Navy, 1962; Al Sarraj, 1990). A schedule that
uses the line of balance method typically emphasizes spatial information in addition to the
time information. The line of balance is especially useful for repetitive work that is performed in multiple locations on the project site. For example, multiple floors of high-rise
buildings, and highway pavement.
Bridge loss modeling. Schedules for repair of highway bridges are used in various loss
estimation methodologies. Mackie et al. (2008c) uses estimated schedules for bridge repair

91
as data for repair duration and repair effort on specific reinforced concrete bridges. Werner
et al. (2006) also estimates scheduling data in order to determine probabilistic bridge downtime throughout a highway network after earthquakes.

4.3 Application to highway bridge loss estimation


This example of highway bridge loss estimation methodology Mackie et al. (2008a)
illustrates the importance of using the communication patterns in describing preliminary
information. Understanding pattern usage facilitates conversation with the various participants involved in collecting data to support this methodology: structural engineers, bridge
estimators, and Caltrans decision-makers. Failure to communicate preliminary information
clearly among the participants can lead to confusion and lack of progress. This is important because the implementation of the PBEE methodology required specifying enough
information to estimate future project repairs, yet at a very early stage during preliminary
design.
In some cases, repair estimation can only be done at stages of design where enough
information exists to construct the bridge. But, for bridges of "typical" construction, repair
estimation can be done early in design because many details and sub-structures can be
selected during an immature design state. This is the situation with conventional highway
overpass bridges which have similar details and similar superstructure design (Ketchum
et al., 2004). The opposite situation occurs for unique signature bridges, such as the self-

92
anchoring suspension span of the new San Francisco-Oakland Bay Bridge, which has very
few details that can be known during early stages of design (Bay Bridge Public Information
Office, 2008; Sun et al., 2004).
Understanding the different categories and qualities of design information is also important because much of the information necessary to determine post-earthquake repairs
involves material, spatial, and process information that is not usually part of preliminary
structural design. For example, identifying the materials and processes involved in installing temporary supports and installing column jackets are required for computing the
repair time and repair cost, but are not needed for completing the structural design on the
bridge.
The local linearization repair cost and time (LLRCAT) loss estimation methodology
implementation and its application to bridges are presented in Mackie et al. (2008a). The
example presented here adopts that methodology and focuses on the process used to generate the preliminary design information and on the communication patterns used to express
that design data at the appropriate level of detail. Table 4.2 shows the procedure broken
down into individual steps along with the communication patterns used in each step.
Step 1: Define bridge design information.
One of the challenges to defining the bridge design information for the loss methodology is the fact that this is done at an early stage of design where alternatives are still being
considered and it is not desirable to spend lots of time detailing any particular option in order to reduce rework. So, this challenge of defining bridge information is to define enough

93

Table 4.2: Loss estimation procedure for reinforced concrete bridges.


#

Step description

1 Define bridge design information

Communication Patterns applied


Spreadsheet (4.2.4),
Sketch (4.2.2)

Define performance groups and damage states

3 Compute EDPs from a mechanics model

Narrative (4.2.1)
Mechanics Model (4.2.5),
Spreadsheet (4.2.4)

Create damage scenarios

Sketch (4.2.2), Annotation (4.2.3)

Estimate repair cost and repair schedule for

Spreadsheet (4.2.4),

damage scenarios

Schedule (4.2.8)

Obtain repair cost and repair effort (production

Spreadsheet (4.2.4)

rate) for each repair quantity


7

Run LLRCAT program to produce results

Spreadsheet (4.2.4)

94
of the bridge so that losses can be estimated, but not to require too much such that the information would not be available this early in the design. This means that the information
presented cannot be at the level of precision required for construction or purchasing, but
should be at the level of precision needed for defining a mechanics model and ensuring that
the structure will be able to perform its basic functions.
The bridge information at this stage is defined using two different communication patterns: the Spreadsheet (4.2.4) and the Sketch (4.2.2). A spreadsheet gives the numeric
values needed for the bridge. Most of the values presented in this spreadsheet refer to the
external dimensions of the bridge superstructure, roadway deck, abutments, and foundations. This allows the computation of volumes and weights so that the amounts of quantities can be estimated based on size. Some basic material properties must also be defined
such as unit weights for steel and concrete, and the compressive strength of concrete f'c,
and yield strength of reinforcing steel fy.
Only a few pieces of information are needed beyond the basic dimensions. This means
that the rest of the procedure can generally continue even before the state of the design
reaches the level of maturity needed for construction. So, the loss estimation method can be
performed at early stages of design, which is when performance information can have the
most influence on the final design without requiring substantial rework. Basic dimensions,
sizes, and material counts can be estimated based on previous designs and preliminary
design tools.

95
Step 2: Define performance groups and damage states.
A performance group (PG) is a group of one or more bridge parts that contribute significantly to repair-level decisions; are repaired independently from other performance groups;
and, have damage which can be characterized by an EDP from the mechanics model. Performance groups are not necessarily the same as load-resisting structural elements, and can
include "non-structural" components because they also suffer damage and contribute to
repair costs.
The unit of analysis needed for design is often different from that which is needed
for repairs. Grouping bridge parts into a performance group allows for groupings that are
specifically suited for analyzing damage and repair. In the analysis for design of seat-type
abutments, several distinct functional components: shear keys to limit transverse motion,
back walls to limit longitudinal motion, and bearings to connect superstructure and abutment. Each components must have its own material properties and force-displacement
behavior. But, understanding the repair of seat-type abutments, such distinctions are unnecessary, and it becomes necessary to treat all these parts as a whole system. Bearings
need to be replaced when their strain capacity is exceeded, and this only occurs when there
is damage to the shear key or back wall which allowed this excess displacement to occur at
all. Thus, it is better to handle these repairs as always linked to one another.
These different units of analysis needed to characterize repair are the performance
groups for the structure. The performance groups are connected to the mechanics model of
the structure via the EDPs that characterize the damage of the performance group.

96
Steps involved in defining performance groups:
(a) Use best judgment to enumerate the different types of possible post-earthquake damage and repairs. Data on damage and repairs from previous earthquakes may be unavailable because new designs may have never experienced a major earthquake, and
bridge replacement may have been preferred over bridge component repair. Group
discrete repairs if they are performed together.
(b) Create the PGs by grouping the physical portions of the bridge where these repairs
are located.
(c) Choose the EDPs that will characterize the damage of these individual performance
groups.
(d) Define one or more damage states (DSn) for each PG by ranking the damage to
each PG in order of ascending severity. Also define another damage state (DS)
corresponding to the maximum possible severity, which can be equal to the highest.
(e) Use prior data on earthquake damage to assign EDP values corresponding to the
different damage states.
(f) Check for overlapping repair methods among the performance groups. If there are
overlaps, then either consolidate multiple performance groups, or split up into smaller
groups so that there is no longer any overlap.
Once the performance groups with their corresponding repair methods and damage
states are defined, they need to be described using the appropriate level of detail. This is
a thorny issue because we only have preliminary design data to deal with and are working

97
with a design that does not have enough maturity to be constructed. Because of this relative
lack of maturity in the design, it is not possible to describe things at a Describe the PGs
using a Narrative (4.2.1), and input the numerical data into the Spreadsheets (4.2.4) of the
methodology.
Step 3: Compute EDPs from a mechanics model.
After the dimensions of the bridge are determined and the EDP responses are known,
then a mechanics model can be created that accurately represents the bridge.
The information needed to create the mechanics model might be different than that
needed for the bridge design information. This is because the model requires selection of
such things as material models and numerical analysis routines, which are necessary for
computing the model, but are unnecessary for constructing the bridge. In the creation of
the mechanics model, there may be the need to create additional Abstract (5.1.1) variables
to represent the needed information.
The mechanics model must be very specific to the individual structure, because differences in site conditions and bridge geometry might not affect the definition of performance
groups and repair methods, but could have an impact on bridge response during earthquakes.
The mechanics model should be analyzed with a suite of ground motions for the site at
different levels of the chosen intensity measure. This will produce a range of data that can
be used to interpolate the performance of the bridge as a function of earthquake intensity.
For each ground motion, the EDPs for each performance must be recorded. These values

98
will produce a Spreadsheet containing all of the numerical data in a summary form. The
time histories are not needed, only the value of the corresponding EDP. The types of EDPs
could be the maximum values throughout the time history (e.g. maximum column drift),
or the value at the end of the time history (e.g. residual column drift).
The procedure for running a bridge model and selecting analysis methods and IMs are
described elsewhere (Mackie, 2004; Mackie et al., 2008a).
Step 4: Create damage scenarios.
A damage scenario is a correlated assembly of performance group damage states. The
scenarios are to be used for estimating the cost and durations of repair. The scenarios
should also be representative of the different damage states and should cover most if not
all of the different damage states so that it will be possible to understand the cost and time
involved in all the different repair methods.
The scenarios do not necessarily represent real instances of earthquake damage, but
are intended to reveal dependencies in the different repair methods and provide a realistic
indication of how much various repairs might cost in combination with other repairs.
The steps for creating damage scenarios are:
(a) Decide on the number of damage scenarios to create based on the bridge design
information and the nature of the expected bridge damage.
(b) For each damage scenario, group the damage states for each performance group together in a way that it is feasible for all the chosen damage states to be achieved
together in the bridge.

99
(c) Decide on the repair methods to be used in each damage state.
(d) Create drawings with Annotations (5.1.1) showing the extent of the bridge repairs to
perform, and the quantities of each repair item.
Good damage scenarios represent the correlations among the different damage states in
each performance group. The mechanics model provides these correlations through deformation compatibility for the deformation-based EDPs, and force equilibrium for the forcebased EDPs. For example, good scenarios will have similar damage states for elements that
undergo similar displacement. A bad scenario would have vastly different damage states
for those same elements. Intuition and judgment are needed to design the repair methods
and repair quantities in a manner that suits these correlations.
Assigning repair quantities can be challenging because specific details are not yet fully
available early in the design process. Usually, full estimation of repairs does not happen
until after making a commitment to details down to the level of individual rebar. But, this
information on repair is most useful during these earlier stages because then the analysis
can have greater influence on overall project delivery. This challenge can be overcome by
using case study repair data from similar bridges after previous major earthquakes, laboratory testing data, or mechanics model results.
The most important part of this step is the creation of drawings that communicate the
damage scenarios in a clear and spatial way. This enables the estimation step to proceed,
and provides enough information so that the damage scenarios can be critiqued and checked
by the bridge engineer who estimates the scenario repairs.

100
Step 5: Estimate repair cost and repair schedule for damage scenarios.
Each of the damage scenarios should have enough information to estimate the cost and
schedule for the repairs.
The cost estimates should include a total amount for each separate type of item item. It
should provide enough information to determine a unit cost for each repair item. Obtaining
this unit cost allows the information to be reused for all the different damage states that
might be encountered in the structure.
The estimating process also serves as a check on whether the estimating quantities for
the repair methods were correct and whether there is enough information in the structure
design in order to estimate. This is a difficult issue because there might be insufficient
information to estimate quantities if the design lacks sufficient maturity. The estimating
process will reveal any missing quantities. For example, an item for installation of a steel
jacket column casing, might be missing the associated excavation and backfill needed at
the column base.
The schedule estimate also reveals where certain types of work can be performed concurrently with other types of work. This would affect the repair duration, but not the repair
effort, which is a measure of the number of crew working days required to complete the repair. It is assumed that some durations can be shortened if additional resources are allocated
to a project.
Step 6: Obtain repair cost and duration for each repair quantity.
The unit costs and durations for each repair item can be derived from the estimating data

101
provided in the previous step. The variations in the unit costs between the different scenarios provide some statistical data that can be used to estimate a mean value and dispersion
on the costs and durations.
The method for obtaining unit costs involves:

(a) For each repair quantity line item in the cost estimate, divide the cost by the total
quantity.
(b) Choose the most representative value among the different scenarios to use for the
unit costs in the methodology
(c) Create a Spreadsheet summarizing this information

The method for obtaining durations is more complicated because it involves the summation of several sub-tasks that are reflected in the schedule estimates, but that are not
directly tied into a repair quantity. This methodology is based on using repair quantities as
the interim variable needed to compute the two decision variables of repair cost and repair
time.
The procedure for obtaining production rates for repair effort involves:

(a) Group the schedule activities by the damage states for each performance group.
(b) Sum the durations of time for each activity for each damage scenario
(c) Create a spreadsheet containing the minimum, average, and maximum amount of
effort needed to perform each repair item.

102
Step 7: Run LLRCAT program to produce results.
The final step is to ran the LLRCAT Matlab program that reads all of the design and
model information from Excel spreadsheets and computes the probabilistic repair costs and
times (Mackie et al., 2008a).

4.4 Application to building information modeling


New advances in technology for building information modeling (BIM) open up new
possibilities for work flow in the architecture, engineering, and construction (AEC) industry. Tools for BIM allow the storage and representation of a much greater precision and
range of information than available before. Additionally, the implementation of new data
exchange standards allows for greater electronic interchange of information among different software applications. These advances provide greater opportunity to communicate
design information in a way that is more clear and complete with less errors.
Nevertheless, the limitations of computer models must be understood in order to avoid
the pitfall of poor design communication. Although BIM technologies have their advantages, they also have disadvantages in communicating design information which must be
understood. For example, Day (1997, p. 156) points out a conceptual issue that is encountered when using today's general-purpose BIM tools for communicating spatial design
information:
One of the problems encountered when trying to construct a full 3D model
rather than a set of drawings is that the model lacks the abstraction of the drawings and there is a tendency to keep adding detail, thus making it unwieldy.

103
When working in 3D at the detailed design stage this lack of abstraction can
also lead to frustration as it is impossible to construct many parts of the building because of decisions that have not yet been made. This either results in
time being spent trying to make these decisions, perhaps at a time which is
earlier than appropriate, or the decision is not made leaving the model incomplete. This incompleteness then becomes a problem itself as it has a knock-on
effect on other parts of the building.

This difficulty comes into sharper focus when examining the definition of BIM technology. Eastman et al. (2008, p. 13) defines BIM as "a modeling technology and associated
set of processes to produce, communicate, and analyze building models", which is characterized by the following four properties:
Building components that are represented with intelligent digital representations (objects) that 'know' what they are, and can be associated with computable graphic and data attributes and parametric rules.
Components that include data that describe how they behave, as needed for
analyses and work processes, e.g., take off, specification, and energy analysis.
Consistent and non-redundant data such that changes to component data are
represented in all views of the component.
Coordinated data such that all views of a model are represented in a coordinated way.

Eastman et al. (2008, p. 15-16) also contrast BIM with four aspects of models that are not
BIM: models that contain 3D data only without object attributes, models with no support
for behavior, models composed of combined 2D CAD reference files, models that allow
changes in one view that are not automatically reflected in the other views.
Based on Eastman's definition on what is not BIM, there are some design communication aspects that are not suited for BIM tools. Early in the design process, the object
attributes might not be well known or determined. This is especially true when commu-

104
nicating abstract functional information, which might not yet be related to individual elements or material characteristics. Also, some parts of the structure might not yet have
a well-defined and understood behavior until later in the design process. In these early
stages, the ability to communicate and define information without full object behavior is
essential. This is the reason why tools such as the Narrative (4.2.1), Sketch (4.2.2), and Annotation (4.2.3) patterns are needed for communicating preliminary design information at
the appropriate level of information, without implying too much information prematurely.
Eastman et al. (2008, p. 14) also define parametric BIM objects by their properties:
[Cjonsist of geometric definitions and associated data and rules.
[Gjeometry is integrated non-redundantly, and allows for no inconsistencies. When an object is shown in 3D, the shape cannot be represented
internally redundantly, for example as multiple 2D views. A plan and elevation of a given object must always be consistent. Dimensions cannot
be 'fudged'.
Parametric rules for objects automatically modify associated geometries when inserted into a building model or when changes are made to
associated objects. For example, a door will fit automatically into a wall,
a light switch will automatically locate next to the proper side of the door,
a wall will automatically resizes itself to automatically butt to a ceiling or
roof, etc.
Objects can be defined at different levels of aggregation, so we can define a wall as well as its related components. Objects can be defined and
managed at any number of hierarchy levels. For example, if the weight of
a wall subcomponent changes, the weight of the wall should also change.
Objects rules can identify when a particular change violates object feasibility regarding size, manufacturability, etc.
[Ojbjects have the ability to link to or receive, broadcast or export sets
of attributes, e.g., structural materials, acoustic data, energy data, etc. to
other applications and models.

These properties of BIM objects force the combination of spatial information along
with functional information and the behavior of pieces in the structure. But, this might not

105
be desirable when the designs are immature and there is insufficient design information to
show the spatial relationships. On this concept, Doumont (2002b) comments that "Pictures
are indeed superior for conveying intuitive or global information. By contrast, they do a
poor job of expressing abstract concepts and lack the accuracy that words are endowed
with. Words, in a sense, are worth a thousand pictures." This is why the Narrative pattern
(4.2.1) is important in cases where the design information is not yet rich enough to support
accurate spatial information.
The ability to show different levels of aggregation is good only when detail is available
down to the level of no aggregation. For example, during preliminary design, only groups
of rebar in reinforced concrete might be known, but not the individual bar layouts. A good
tool for communication would indeed allow the different levels of aggregation, but still
allow the use of the less detailed levels without forcing the definition of the more detailed
ones until later in the design process.
These factors indicate specific properties that BIM processes need in order to support
performance-based structural engineering design:
First, the ability to include abstract information that is not specifically tied to spatial elements in the structure. This is important for expressing intensity measure (IM). Structureindependent IMs for earthquakes depend on site conditions alone, and are not related to
any structural objects. Structure-dependent IMs require both site condition information
and general structure properties, such as the fundamental period, T\. The value of T\ is
not tied to a specific part, but is a general property of the whole structure. This is easily

106
accomplished using freeform communication patterns, and should be included within BIM
tools as well.
Second, the ability to express abstract and functional information without requiring
a high degree of spatial precision. Determining engineering demand parameters (EDP)
only requires enough information to define an adequate mechanics model. Often this only
requires preliminary design information for basic structural system behavior, and can be
done before high precision details are available. For example, the preliminary specification
for a lateral force-resisting system might contain only basic abstract quantities representing
structural behavior, such as damping ratio, stiffness, yield strength, and displacement limits.
That kind of specification might not have highly precise geometric rules, but should still be
capable of integration into the building model.
Third, the ability to express a range of possible structure states. Describing damage
measures (DM) requires the ability to model a range of discrete or continuous damage
states. This definition requires consideration of multiple possible states for the structure
after an earthquake. Furthermore, each value of DM will also trigger a different value of
the decision variable (DV) according the severity of the damage. A range of possibilities
must be defined instead of merely the ideal as-built condition of the structure.
Fourth, the ability to define performance groups with freeform information and customized grouping. Expressing damage scenarios and repair methods requires freeform
information because many repairs are specified conceptually rather than with specific details. Customized grouping is needed because repairs within a performance group may

107
involve many elements in addition to the actual piece needing repair. For example, repairing a bridge abutment backwall may require approach slab replacement because of access
requirements.
Fifth, the ability to define probabilistically uncertain quantities. Evaluating the structure using a performance-based earthquake engineering methodology requires definition of
uncertain quantities. IM data cannot be expressed by a single number, but must include
a distribution and parameters, such as median and dispersion. EDP data is also uncertain
depending on the ground motions suite used for analysis. DM data is uncertain because
of variations in the actual as-built condition of the structure. All of these quantities need
proper expression within a BIM process.
Based on these requirements and communication inadequacies of current software tools,
the following improvements are recommended for future BIM software:
1. Integrated tools for freeform sketching and annotation.
Providing a specific function for creating freeform annotations inside a working
model view would overcome the current barrier of making annotations in a separate drawing tool or window. A feature for taking a model snapshot and scribbling
on it would allow the communication of conceptual information and avoid the tendency to allow software programs to generate the views without considering how
other project participants will understand it. The snapshot freezes a model view and
prevents unintended changes to the sketch. This feature could be combined with the
hyperlinks feature noted below.

108
2. Hyperlinks that connect model entities to freeform descriptions.
Hyperlinking from model entities to freeform narratives and spreadsheets would
overcome the inability to link BIM objects to freeform descriptions. This would
allow the expression of abstract and functional information without requiring a high
degree of spatial precision, because such information could be linked to a volume
or aggregation inside the model instead of a single specific piece within the model.
Linking a 3D model-based visualization of the structure to spreadsheet cells would
also reduce the amount of time needed for spreadsheet data entry. For example, the
geometric spreadsheet data needed for the LLRCAT methodology would be much
easier to navigate and understand if linked to a 3D model view.
3. Present conceptual mechanics models and sketches as another view layer in 2D or
3D space.
This feature would allow different conceptual presentations of the structure to be presented together. For example, a rendered model view could be presented atop a hand
sketch of the same location to communicate design intent in a freeform manner. Or, a
line element mechanics model drawing could also be displayed to illustrate how the
different components act together to perform their intended structural function. This
would violate the principle of non-redundant geometry and composition of multiple
2D references, which Eastman et al. (2008) use in their definition of BIM. But, it is
necessary for ensuring good design communication.
4. Capture design intent at the moment of model object creation.

109
A key ingredient in the design process is not only the final material and spatial specification of the building, but also the functional reasons for their selection and design.
Presenting an option for the designer to enter freeform information about why a particular choice was made would enable automatic annotation of this design intent. For
example, this feature would allow the designer to note that a steel floor beam must
be a W14 shape because of ventilation duct size. This additional metadata could be
shown back to remind the designer if the shape were later changed. This feature
would help overcome the lack of support for abstract design information in current
BIM software.
5. Assign probabilistic uncertainties to numeric values inside the model.
This feature would overcome the inadequacy of using a purely deterministic model
for evaluating probabilistic performance. Performance-based earthquake engineering methods are inherently probabilistic, and providing this information would aid
performance evaluation. This might be implemented by adding fields for naming a
probability distribution and for the values of the distribution's parameters. This additional information could be used for probabilistic structural analysis, and ignored
for other purposes where the median value could be used deterministically instead.
The advances in BIM technology are indeed exciting for the field of structural design.
However, caution is needed in order to avoid neglecting the strengths of time-honored design communication principles and methods. Implementing these suggested features would
improve the ability of BIM software to use the patterns of good design communication.

110

Summary

What I have done in this chapter:

1. Provided patterns for communicating design information at the appropriate abstract


and concrete levels of detail.
2. Illustrated the necessity of annotation and text in order to describe building model
attributes. Annotation communicates the appropriate amount of conceptual information, while the building model (and graphical views) communicate attribute data.
3. Provided an example for communicating abstract performance criteria without overspecifying the level of detail required to check it.
4. Commented on the use of BIM technology for communicating design and suggested
recommendations for improved design communication.

Ill

Chapter 5
Analysis and decision patterns
The patterns of analysis and decision for structural design involve approaches for determining the information about the structure. The patterns for analysis describe ways that
additional information about the structure are created. This analysis information is not
necessarily required for constructing the actual structure but are required for the design
process. The Abstraction pattern describes the creation of abstract quantities in order to
facilitate the design process. The Aggregation pattern describes the concept of grouping
for physical items, and the Agglomeration pattern describes the concept of grouping for abstract items that relate to function rather than material composition. The Updating pattern
shows an example of how design information is changed and updated. The Engineering and
Social Performance patterns show the formation of criteria for judging whether a design is
correct, and provide measurements for choosing one option among many. Engineering Performance mainly focuses on the product, while Social Performance focuses on the people

112
and process involved in making the product.
The patterns for decision making describe ways of choosing more detailed spatial and
material information based on abstract design information and the goal of fulfilling the
functional requirements. They describe different ways of arriving at more mature design
information, and use the information from analysis results to make decisions. The Pick Any
pattern simply chooses any permissible item among a set of feasible choices. The Minimum
and Maximum patterns describe methods of selection based on the numerical value of a
single parameter, assuming there are no other tradeoffs in the selection. The Optimization
pattern provides examples of how to make decisions when there are tradeoffs and multiple,
competing criteria involved in making the best design choice. The Prescription pattern is for
decisions that are governed by a set of strict rules rather than intuition. The Set Exploration
pattern is a description of the set-based design method and illustrates the ability to conceive
of novel ways to make decisions based on the same analysis capabilities.
The analysis and design decision patterns proposed here address the common relationships between the design decisions to be made and the qualities of the variables at hand. It
will be shown that under some circumstances, the appropriate tools for solving the design
problem are rather different. The solution could require a detailed engineering calculation,
a conversation between the structural engineering and the general contractor, or a lookup
to a table inside the building code. The proper identification of these different patterns will
help in the understanding of how best to approach the problem in a manner that promotes
innovation.

113
This chapter does not emphasize the actual methods for doing structural analysis itself,
but rather is focused on the conceptual content needed to in the analysis process. Patterns
for executing and running mechanics models are described in Heng and Mackie (2008),
where the design pattern concept is used to to identify best practices in object-oriented
finite element software design. The design patterns for analysis here are not specific to
finite element methods, but rather reflect the kinds of analysis results that are required to
make design decisions.
The names of these patterns, once adopted, can be used as a metalanguage to describe
design intent when using the pattern. Being able to label how a decision was made and
what type of analysis was carried out provides the rational for the information. Gamma
et al. (1995, p. 3) note that using pattern names "increases our design vocabulary", makes
it easier to discuss and document design, simplifies communication about trade-offs, and
allows people to "design at a higher level of abstraction". This discussion of abstract design
enables better design of the design process itself.

5.1 Patterns for analysis


5.1.1 Abstraction pattern
Intent
Create a variable that is itself non-physical in order to make general analysis feasible.
Convert the real spatial and material information into another form to facilitate thinking.

114
Also known as
Mechanics model variable, simplification, modeling assumption.

Assumed background
Foundational + Materials science + Mechanics + Problem recognition and solving +
Design, knowledge of dynamics of structures (Chopra, 2007), and knowledge of basic
earthquake engineering concepts.

Motivation
In mechanics models, it is sometimes necessary to create objects that are not directly
linked to real spatial or material design information, but that are useful for the process
of thinking. These kinds of abstractions allow mathematical manipulation and intuitive
physical insight into a design problem, yet without dealing with how these abstractions
must be instantiated as real, physical objects. Sometimes these abstractions are created out
of convenience as simplifications because the exact nature of the real physical phenomenon
is too complicated to capture. In these cases, only some essential characteristics might be
necessary to express the phenomenon in a meaningful mechanics model.

Applicability
The Abstraction pattern is useful in situations where:
1. A mathematical analysis procedure can be converted into a form similar to something

115
more basic and familiar, such as simplifying analysis of a composite material into an
equivalent homogeneous material.
2. Quantities inside mechanics-models occur frequently enough to be named and ascribed physical characteristics, such as with factors and coefficients that occur frequently in design equations.
3. The actual physical phenomena are too difficult to model exactly, but can be approximated using a simple model to achieve acceptable results, such as with using
simplified material models which are abstract entities.

Structure
The abstraction pattern starts with something that is difficult to handle in its original
form and transforms it into another things that is more familiar, or more general. They are
discussed as though they were real entities, although they are in reality concepts and not
necessarily directly based on physical objects.

Consequences and implementation


Abstractions are the objects that are necessary for design, but unnecessary for for construction because they are non-physical. This pattern allows a high level of thinking and
analysis to characterize the essence of an engineering problem at the appropriate level of
detail, avoiding the need to specify details prematurely.
The potential pitfall of creating abstractions is that they require careful judgment to be

116
used correctly. A lack of judgment might lead to false results based on abstract reasoning.
For example, Ferguson (1992, p. 183) laments that "nearly all engineering failures result
from faulty judgments rather than faulty calculations". A design procedure that consists of
abstractions without giving due consideration to the realities of scale, form, and proportion
are more susceptible to lapses in judgment than one that is based on the "subtleties of real
world engineering" (Ferguson, 1992, p. 169). It is necessary to deal with abstractions when
creation conceptual and mechanical models, and good judgment is needed to complete
design beyond the level of precision where the abstractions are applicable.

Known uses
Modulus ratio, n, in transformed section analysis. For reinforced concrete and steel
composites made of materials with different elastic moduli E, use a single value of E
and modify the area A. When calculating an engineering quantity using multiple different
kinds of materials, it can become confusing to deal with each material's properties separately. Using a "transformed section" allows one to perform all the computations using
only one material's properties by using a transformation on the area. In reinforced concrete
elements (MacGregor and Wight, 2005, p. 388-389) and steel composite design (Segui,
2007, p. 554555), the area of the transformed section is usually computed by scaling the
steel area by the modulus ratio n = Es/Ec.
Material models. Assuming a particular form of material model, such as a linear-elastic,
perfectly-plastic stress-strain behavior. This abstraction is based on some physical proper-

117
ties, but is itself simply a mental construction because such materials do not actually exist
and cannot be created. In such a material model, the parameters Gy and E are actually
abstractions that permit analysis and thinking.
r factor in structural dynamics. The factor Tn = s/Mn is a quantity that occurs
frequently in dynamic analysis of structures (Chopra, 2007). It is a quantity that is not
directly physically measured or that is necessary for construction. Rather, this quantity is
an artifact of the equations of motion needed to analyze dynamic structural behavior. It is
useful for the purposes of design and checking engineering performance, but will not help
when evaluating for social performance or construction.
Euler-Bernoulli beams. The Euler-Bernoulli beam theory assumes that plane sections
remain plane. This assumption greatly simplifies analysis, but neglects the effects of shear
deformation and warping. Using this beam theory creates abstractions that approximate
actual beam behavior.
PBEE Variables. The variables in the PEER performance-based earthquake engineering
(PBEE) methodology are examples of abstractions. The intensity measure (IM), engineering demand parameter (EDP), damage measure (DM), and decision variable (DV) represent
concepts for general analysis that can be realized using different physical information.
The IM could be structure-independent, based on the physical earthquake phenomena
alone, such as peak ground acceleration or peak ground velocity. Or, the IM could be
structure-dependent, such as the spectral acceleration at the fundamental period, Sa(7i).
The EDP could be based on forces, displacements, or ratios at their maximum, minimum,

118
or residual values. The DM could reflect a smooth continuum of increasing damage, or be
based on discrete damage states. The DV could be based on dollars, downtime, repair time,
human safety, or other concerns.

5.1.2 Aggregation pattern


Intent
Group similar items by their material or spatial information, and treat conceptually as a
single item. Simplifies a conceptual or mechanics model by reducing the number of pieces.

Assumed background
Foundational + Materials science + Mechanics + Problem recognition and solving +
Design + Communication.

Motivation
In the design of structural elements, the behavior of a single piece of material is not
as important as long as the entire group of material can be assumed to work together.
For example, inside a prestressing cable, the individual strands are usually not considered
separately for design as long as the entire cable's behavior can be understood. Similarly, in
reinforced concrete design, the behavior of each individual bar is usually not considered as
long as the the total amount of steel functions together as a unit.

119
Applicability
Use the Aggregation pattern when many small items perform together as a group and
have similar behavior; individual behavior of individual items is not needed, and only the
overall group behavior is needed; and the material and functional information is important,
but highly precise spatial information is not. The Aggregation pattern is not applicable in
design situations where the spatial location of each individual rebar were significant, as
might be the design of a highly congested element.

Structure
The Aggregation pattern a conceptual device that helps eliminate unnecessary complexity during the design process.

Consequences and implementation


This pattern provides the ability to compute performance without requiring spatial
specificity. For example, the reinforcement ratio p combined with a quantity As and depth
to steel centroid d allow the consideration of rebar behavior in a reinforced concrete element without needing to specify a particular bar layout. These quantities support set-based
design by allowing engineer to work with a set of acceptable rebar layout options by only
dealing with the quality needed for input into the mechanical model. Aggregation uses
abstract parameters in order to represent a group of different objects inside a mechanical
model.

120
Aggregation is a useful tool for specifying information at the correct level of precision.
In the design of reinforced concrete, specifying the exact location of each individual bar to
a high degree of precision might be unnecessary to achieve structural performance.
Aggregation might be disadvantageous when the exact spatial location is needed for
each individual piece. For example, aggregating multiple rebars together in a beam might
be useful for performing section analysis, but would make it difficult to check beam-column
joint congestion.

Known uses
Area of steel, rebar groups. The area of steel As in reinforced concrete design is an example of aggregation. The individual reinforcing bars are treated as a single object in cross
section for checking certain abstract quantities inside a mechanics model. For example,
checking the nominal moment capacity requires the value of As and steel centroid location,
but not the precise spatial arrangement of the individual bars.
Rebar can also be defined by groups instead of individually. For example, a design
could specify the number of bars, bar size, and cover location (e.g. "5-#8 bar @ 2" cover"),
while leaving the exact location of each individual bar up to the fabricator, detailer, and
placer. This works well as long as there is no ambiguity in the specification and the design
is complete.
Reinforcement ratio. The reinforcement ratio p is another type of aggregation based on
the rebars inside a reinforced concrete element. The reinforcement ratio is computed as the

121
ratio between area of steel As and the concrete section gross area Ag. ACI 318, Sec. 10.3.2
defines three different reinforcement ratios, p for tension steel, p' for compression steel, and
Ph as the reinforcement ratio producing balanced strain conditions where the steel yields at
the same time that concrete crushes.
The value of p is important for checking certain conditions in the mechanics model.
It is used to verify that steel yields before concrete crushes for ductile behavior in flexure
elements, and can also be used to estimate rebar congestion for placing difficulty during
construction.
Prestressing strands. The most common type of prestressing strands are seven-wire
strands in a twisted bundle (Naaman, 2004, p. 46-49). Although the strand is made up of
seven individual wires, the entire strand is treated as a single item for structural analysis.
The tension force is assumed to be distributed across the entire area of the each wire, and
acts through the strand center.
Bolt groups. Groups of bolts in steel connection design can be treated for design as a
single reaction force on a member with a possible eccentricity (Segui, 2007, p. 443-^46).
Individual bolts need to be considered for limit states involving the bolts and the holes or
slots. But, once the connection is known to be designed properly, the individual bolts are
usually not necessary to consider to check other measures of engineering performance.

122

5.1.3 Agglomeration pattern


Intent
Group different types of items in a structure by their function in order to facilitate
analysis.

Assumed background
Foundational + Materials science + Mechanics + Problem recognition and solving +
Design + Communication.

Motivation
For structural analysis, the behavior and function of structural elements can be more
important than their actual material composition and exact spatial extent. Agglomeration
is the concept that deals with the creation of conceptual items such as reinforced concrete
"beams" that are physically composites of concrete and rebar intended to resist loads in
flexure. The beam concept itself is an agglomeration of several other materials in a particular spatial arrangement.
This concept is helpful because an agglomeration can be modeled using simpler methods than considering its real composition. For example, a reinforced concrete beam's nonlinear behavior can be characterized in a mechanics model using a M-0 moment-curvature
relationship instead of modeling the individual rebars and concrete matrix. The beam's behavior requires knowledge of the concrete and steel parts, but once agglomerated, knowl-

123
edge of neither is necessary to characterize the component in a model.

Applicability
Use functional agglomeration when there are several different components perform the
same function in the structure, and when accomplishing a particular structural function
requires multiple components that work together. This is not applicable when there is no
related functional information between the different structure components.

Structure
The agglomeration pattern involves grouping different kinds of structure parts by their
function. The process requires judgment and understanding how load-resisting systems
operate and general principles of building form and basic subsystems. This pattern permits
the design strategy described by Lin and Stotesbury (1988, p. 18):
The strategy is to emphasize only the more primary levels of functional
and aesthetic and economic requirements until a promising total system design
has emerged. In this way, consideration of the more detailed specifics of each
entity (the secondary-level issues) is postponed so that they may be influenced
by the total-system context. Such an approach enables the designer to identify
overall issues with clarity and to deal with the design of each spatial subsystem
in a hierarchic fashion.
But this is not to say that the designer should ignore the importance of
structural specifics at the formative stages of his spatial thinking. Indeed, if
the designer is experienced, he or she knows that the ultimate provision of
both overall and specific means for physical integrity is essential. But he or
she will need to have a suitable way of discriminating and then dealing with
only the overall issues of structural design at the formative stages of design
thinking. When this is not the case, it is not surprising that structural (and
other technical) concerns will be more or less overlooked at schematic levels.

124
Consequences and implementation
Agglomeration focuses the design around functional information, which describes how
different components work together. This type of grouping is useful in structural system
design when the interaction of all the components need to be analyzed together. Agglomeration is also helpful for the creation of simplified mechanics models to capture the behavior
of a structure. The elements in a simple finite element model are the agglomerated components of the structure.
Overlooking the connections between agglomerated components and critical details inside them are possible pitfalls to using this pattern in analysis. For example, looking mainly
at the function of a beam and column might cause a designer to overlook challenging details involved in the construction of the beam-column joint. Wisdom will dictate how to
tradeoff using agglomeration to simplify analysis and when to look at the real components
when the details become necessary.

Known uses
Performance groups. Performance groups are composed of several different pieces of
the structure, but are combined in a way that their behavior are treated as a single unit
(Mackie et al., 2008a). The groups represent parts of a structure that are repaired together
after an earthquake and whose structural performance can be characterized by a single
engineering demand parameter. The example in Sec. 4.3 defines performance groups and
their contents.

125
Load paths. Analyzing load paths within a building uses agglomeration. Load path
analysis checks for a complete path for loads to propagate through the structure down to
the foundation. Only the function of the structural components is important for checking
the load path, and the actual material used in the building components is often not essential
as long as components are properly designed.
Elements in mechanics models. Mechanics models often use some degree of agglomeration in order to simplify the analysis. A reinforced concrete beam might be simply treated
as an elastic beam line element for some analysis. In this case, the behavior of individual rebars do not need to be considered, only understood that they exist and perform their
intended function.
But, whenever the individual behavior is needed, the agglomeration must be broken in
order to complete analysis. This may happen with non-linear behavior where a fiber model
is required to separate the material behavior of rebar and concrete, or if the section warps
and the assumptions of the Euler-Bernoulli beam theory become invalid.

5.1.4

Updating pattern

Intent
Use when an early initial guess needs to be updated after receiving additional design
information.

126
Also known as
Recompute, refine, "sharpen the pencil".

Assumed background
Foundational + Problem recognition and solving + Design + Risk and uncertainty.

Motivation
An initial guess might need to be updated once more information comes in. Design
information with less maturity will change after the designer determines more information.
Sometimes an initial estimate on a value is needed in cases where the material variables
of the object being designed are themselves variables in the problem solving method. In
the design of beams, the beam's self weight contributes to the load on the beam. The
estimated load is needed to estimate the beam size, then the beam size is needed to update
the estimated load. The updating pattern occurs when there is interdependency in a design
process (Steward, 1981; Browning, 2001).

Applicability
Use the Updating pattern:

1. When there are interdependencies among the design variables, and one variable has
to be assumed in order to get the problem solving algorithm started. Overcomes

127
Start with initial
values of variables

Use XQ

to determine
values of y

XKiXQ

Use y to update the


values of XTHX\

Complete analysis
using x\ and y\

Figure 5.1: Updating pattern.


the "chicken and egg" problem when beginning to make design choices based on
information from mechanics models.
2. The level of detail and precision increases during the stages of the problem solving
algorithm. A lower level of detail and precision can be used in the beginning, then
updated to a higher level later in the process.
3. When additional data will become available in the future, and you want to use that
data to influence decisions.

Structure
The general structure of the Updating pattern is illustrated in Figure 5.1. Initial values
of the quantities x ss XQ are assumed in order to compute other quantities y\, which are then
used to update the initial guess to a more accurate value x ~ x\.

128
Consequences and implementation
The Updating pattern is useful when it is not possible to continue analysis with abstractions. The abstract information needs to be instantiated into combinations of real material
and spatial information, closer to the information needed for construction. This allows the
designer to start with a rough approximation in order to get started on a design task. But,
this process will require design information to be revisited again, requiring some iteration
and possibly rework. Implementing this may be difficult if a high degree of precision is
needed in the initial guess and insufficient information is available to guide the selection,
similar to the problem of "starvation" noted by Terwiesch et al. (2002).

Known uses
Estimating dead load. When sizing structural elements that carry their own self-weight,
an initial value of the element's self-weight must be determined in order to determine the
loading on the element. Once the size of the element is chosen, then the actual self-weight
can be added to the original loading to check if the element still meets the design criteria.
Preliminary steel frame design. Lee and Goel (2001) updating moments after choosing beam and column shapes. Before choosing beams and columns, design forces Fi are
estimated based on estimates of the fundamental period, yield drift, and a selected target
drift. These design forces are used to determine beam sizes using the AISC-LRFD method.
After the beam sizes are chosen, the design forces are updated F-m for the final selection of
beams and columns.

129
Structural health monitoring. Data collected from the actual performance of a building
can be used to update estimates of performance based on variables used in the design process. This process is often referred to as structural health monitoring, where the condition
of the structure, "health", is monitored using sensors or human observations. The development of sensors and methods for performing updating based on sensor data is a topic
of much current research given the recent innovations in sensor technology and wireless
communication (Lynch and Loh, 2006). Methods of using sensor data to update bridge
performance within the PEER PBEE framework have also been proposed by Wong et al.
(2006); Wong and Stojadinovic (2006).

5.1.5 Engineering Performance pattern


Intent
Measure structural design performance using the results of a mechanics model related
to the stress and strain, force and displacement of a structural model.

Also known as
Structural performance, engineering demand parameter (EDP), limit state, mechanics
model results.

130
Assumed background
Foundational + Mechanics + Materials science + Problem recognition and solving +
Design + Sustainability + Risk and uncertainty + Contemporary issues + Project management + Communication + Public policy, and comprehension of performance-based earthquake engineering (Cornell and Krawinkler, 2000; Moehle et al., 2005).

Motivation
Engineering performance provides a way of transforming qualitative requirements of
building function into quantitative requirements that can be checked using engineering
tools. Engineering performance measures how well the materials in the structure are able to
perform their intended function to resist loads and provide serviceability. These measures
generally come from a mechanics-based model which gives information about stresses,
strains, forces, and displacements in the structure. Other analysis outcomes such as vibrations and aerodynamic effects are also important for characterization depending on the
unique aspects of each structure.

Applicability
This pattern is applicable when the performance requirements are based on the structure's intrinsic properties. This encompasses the physical behavior of the structure and
its response to loads. This pattern converts qualitative and descriptive requirements into a
quantitative form that can be evaluated using mechanics-based models and mathematical

131
methods.
This pattern is not applicable when the requirements are not fulfilled through the material behavior of the structural components. In those cases, the descriptive requirements
cannot be translated into a quantitative form, and therefore a different type of performance
measure must be used to evaluate.

Structure
The Engineering Performance pattern involves most of the quantities related to structural engineering analysis. This includes such things as maximum material capacities like
plastic moment capacity Mp, global displacement measurements like drift ratios, and local
displacement measurements like maximum element deflections.
The pattern works by taking some functional requirement and translating that requirement into quantitative requirements that can be checked with an analysis method. For
example, a residential balcony that has cantilever beams needs avoid damage to finishes
and remain level enough for people to use it without feeling unsafe. This might translate
into a maximum deflection at the balcony end, which provides a minimum required beam
stiffness. The deflection and stiffness are the Engineering Performance parameters that can
be checked in a mechanics model in order to satisfy the requirements.

132
Consequences and implementation
Implementation depends on the type of performance measure considered. The concept
of engineering performance typically refers to quantities that can be measured and represented in a mechanics model of a structure.

Known uses
Almost all engineering analysis procedures use this pattern tacitly by aiming to produce
results for evaluation and comparison to a standard of acceptable performance.
The PEER PBEE methodology links specifically links engineering performance to
functional requirements by using the total probability theorem. This is accomplished by
defining functions that related earthquake hazard, structural performance, and the possible
consequences (damage, downtime, dollar cost) to the parameters that drive decisions about
performance requirements. The engineering demand parameter (EDP) in this methodology
is an explicit example of the Engineering Performance pattern.

5.1.6

Social Performance pattern

Intent
Measure structural design performance using information about social conditions.

Also known as
Decision variables, non-engineering factors, "soft" issues.

133
Assumed background
Foundational + Mechanics + Materials science + Problem recognition and solving +
Design + Sustainability + Risk and uncertainty + Contemporary issues + Project management + Communication + Public policy + Teamwork, and comprehension of performancebased earthquake engineering (Cornell and Krawinkler, 2000; Moehle et al., 2005).

Motivation
Mechanics models alone are insufficient to characterize all performance factors for design. Factors such as cost, construction time, construction phasing, material availability,
and estimated post-earthquake repair cost play a role in design decisions, but cannot be
estimated based on mechanics-based structural analysis models alone. They involve factors that are outside the usual realm of mechanics-based analysis. These situations call for
Social Performance criteria that involve this additional information beyond mechanics.

Applicability
This pattern is applicable when the performance requirements are based on external
factors other than the structure's intrinsic properties in as-built state. This includes factors
involving the construction process, cost, time, and issues related to labor and scheduling.
This pattern analyzes performance measures that are not based on mechanics models, but
rather on issues of a social nature that emphasize process information over material and
spatial design information.

134

Design
information

Additional
social or market
information

Social
Performance
measure

Figure 5.2: Social performance.


Structure
The structure of this pattern generally includes design information plus additional social
information and a methodology for combining the two types of information (Figure 5.2).
The additional information is often tacit information based on experience, or particular
industry-specific data. For example, for estimating the cost of a reinforced concrete shear
walls, the additional information could include the prices of rebar at different sizes and
grades, the cost of concrete, the material availability of aggregates, and the optimal dimensions for formwork erection.

Consequences and implementation


Including Social Performance into a design process is difficult because of the need for
additional information that is external to the design itself. The use of tacit, informal information is also a challenge to implementing methods for incorporating Social Performance
into design processes. Factors such as labor productivity are often estimated in a qualitative manner which makes it more difficult to convert to a quantitative value. Much of this
information is not traditionally part of the structural engineering design process. Therefore, implementing this pattern might be very difficult because it forces collaboration and

135
communication in various ways that may only be available in a collaborative environment.
Applying this pattern early in the design process will allow much better decision making
because of greater knowledge about downstream conditions and the largest constraints on
executing the project. Social Performance measures are important for determining project
feasibility and profitability. But, Social performance measures are also more difficult to
evaluate when the design is still immature because there may be insufficient precision to
specify particular products and materials or specific processes for construction. This difficulty can be overcome using case study data from previous projects and general rules
of thumb, such as the suggested rules for limits on concrete reinforcement ratio p to aid
constructibility in ACI 315.

Known uses
The PEER PBEE method uses "decision variables" (DVs) in order to express Social
Performance (Cornell and Krawinkler, 2000). Mackie et al. (2008 a) describe an application of the PEER PBEE framework for computing repair times and repair cost ratios for
post-earthquake damage of reinforced concrete bridges. The design information for the
bridges are based on Ketchum et al. (2004), and OpenSees-based mechanics models provide engineering performance data. The additional information needed to compute social
performance was obtained through multiple interviews with Caltrans bridge engineers and
cost estimators. The mechanics-based and social-based information are combined using
the local-linear repair cost and time methodology (Mackie et al., 2008c).

136
Parrish et al. (2008b) uses this pattern in describing the notion of "value propositions",
and illustrating with labor production rates for different reinforcing steel sizes and bending
types. The additional social-based information is obtained from interviews with rebar fabricators and rebar placers. This information is combined with material and spatial design
information for rebar size and bending type, and with functional information describing the
type of structural element (e.g. wall, column, beam).

5.2 Patterns for decision making


5.2.1 Pick Any pattern
Intent
Choose any option from a set of options, then confirm whether it works.

Also known as
Guess and check, trial and error.

Assumed background
Foundational + Mechanics + Problem recognition and solving + Design.

137
Motivation
With a mechanics model that does not contain abstractions which are able to characterize entire sets of different real design options, it is often necessary to pick any one candidate
solution and check whether it satisfies the design criteria.
This pattern is also useful in situations where the mathematics is too difficult to carry
forward a symbolic solution, and numbers are needed. The Pick Any pattern allows the
selection of values for the design variables, allowing the design process to move forward.

Applicability
The Pick Any pattern is applicable when there is little information available to discern
any differentiating benefit between a set of options. If all the options are the same for all
measures of performance, then it is okay to pick any option.
Often, this pattern is misused when the options have very different performance. Many
decisions in common structural engineering problem solving algorithms implicity use the
Pick Any pattern. But, the only performance criteria evaluated for selection involve structural performance based on abstract variables and a mechanics model. These applications
rarely consider the downstream implications on the rest of the delivery supply chain. For
example, different sizes or grades of rebar might be unavailable, or the selection of a steel
shape might cause clashes with duct work.

138
Structure
The Pick Any pattern requires abstract design information and basic performance criteria that define the design choice to be made (Figure 5.3). The choice is usually to define
more mature or more precise material and spatial information. For example, the choice
might be for the number and size of rebars in a concrete beam given a requirement of
As > 2.00in.2. The criteria for design involve the specification of an Aggregation of steel,
limitations on spacing and dimensions, a required cover, and beam dimensions b and d.
The Pick Any pattern is used to choose any particular instantiation of rebar that fulfills
those requirements.

Consequences and implementation


The Pick Any pattern will lead to a solution if the choice is good and satisfies all the
required performance criteria. If the choice is bad, then it will require iteration to determine
a workable design.
This pattern tends to overlook other options that might be far better than the initial
choice. This pattern also tends to eliminate options which may be better than the initial
option. It does not allow consideration of the downstream consequences.
Implementation is simple, because an arbitrary choice can be made among options that
already satisfy the acceptance criteria. In reality, choices are not completely arbitrary,
because tacit preferences inform a designer's selection.

139

Abstract design
requirements

Determine
performance
criteria for design

Pick Any

Evaluate
performance of
design choice

Satisfies \ N 0
performance
criteria?
Yes

Proceed
with design

Figure 5.3: Pick Any pattern.

140
Known uses
The pick any pattern is used in reinforced concrete design procedures for choosing rebar
layout (MacGregor and Wight, 2005). After the mechanics model constraints on rebar are
selected then any combination satisfying the constraints is selected and used in the rest of
the design process.

5.2.2 Minimum Weight pattern


Intent
Choose the design option that provides the least material weight.

Also known as
Lightest section.

Assumed background
Foundational + Mechanics + Problem recognition and solving + Design.

Motivation
Steel and concrete are priced by weight. So, reducing weight attempts to minimize cost.
But, this does not work when overall cost is controlled by other factors such as installation
labor.

141
A lighter structure might more efficient due to having less material, reduced gravity
loads, and reduced seismic mass. Choosing structural elements that have the least weight
will fulfill these goals as long as the choices meet all the other design criteria. This pattern
most commonly occurs in the selection of steel shapes because the AISC manual provides
design aids for choosing minimum weight and names steel shapes by their weight per foot.

Applicability
The Minimum Weight pattern can be applied in situations where the weights of each
option are easily computed and have a direct influence on the design decision. This would
be the most applicable in the selection of structural members for framing or support. In the
selection of rebar layouts, selection by weight is probably not useful because all choices
will have approximately the same weight since they must all provide a similar area of steel.
For rebar, congestion often controls cost more than material weight.
Using this pattern is better than the Pick Any pattern because it provides a more intelligent basis for determining the material and spatial information needed to complete the
design.

Structure
The Minimum Weight pattern picks the design option that has the least weight and
fulfills all other structural criteria. This works best when there is a catalog from which
to choose, or some clearly understood set of options. When the design options are more

142

/ ,

Table 3-2 (continued)

W Shapes
X

Selection by Zx

**

v\

in.3

ASD

W36x150
W27x178
W33x152
W24x192
W18x234"
W14x283h
W12x305ft
W21x201
W27X161

598
581
570
559
559
549
542
537
530
515

1490
1450
1420
1390
1390
1370
1350
1340
1320
1280

W33x141
W24x176

514
511

1280
1270

Shape

W40x149v

/> = so ksi

kip-ft
H

%Mpx

kip-ft

kip-ft

LRH'I"
2240
2 , 8U
2140
2100

..! '1

2011)
1 VJl.

m*
kip-ft

kips

kips

L,

'*

ASD

LRFD

ASD

LRFD

ft

ft

898
880
882
851
858
814
802
760
805
800

1350

58.U
M 8
.,.'7
4fl.
28 U

121u
".cU'J

38.6
34.5
21.7
32.0
18.7
10.8
5.53
4 66
14.6
20.8

8.09
8.72
11.5
8.72
10.8
10.1
14.7
12.1
10.7
11,4

782
786

1180
1180

30.4
18.3

45.8

8.58
10.7

' 'M.'j

1930

!.'(:
; <
' % .

12?C
1200
1.'.l)

fif

:/
7 00
.'

>>

31 3

"t/Or

kips

kips

in. 4

ASD

LRFD

23.5
25.2
36.3
25.7
39.6
61.5
114
137
46.1
34.7

9800
9040
7020
8160
6260
4900
3840
3550
5310
6310

432
448
403
425
413
489
432
530
419
364

650
672
605
638
619

25.0
37.4

7450
5680

403
379

604
568

733
648
796
629
546

Figure 5.4: Minimum weight design aid sorts steel W-shapes by weight and performance
(AISC Manual, Sec. 3).
freeform, more information might be required.
This pattern is best illustrated using the selection of a steel beam from the catalog of
standard AISC W-shapes. The AISC Manual, Sec. 3 contains tables of steel shapes sorted
by required plastic modulus Zx and arranged in subgroups by weight. The least weight
section is positioned at the top of each subgroup and indicated by bold text.

Consequences and implementation


Implementation of this pattern is typically straightforward. Determining total weight is
usually a simple computation that can be done using basic tools. Design aids for choosing by minimum weight are readily available for structural steel. The AISC Manual pro-

143
vides tables for choosing least weight W-sections by minimum required engineering performance, measured by either plastic strength Zx (Figure 5.4) or stiffness /.
When applied haphazardly, this pattern could lead to excessive variations in the total
design. For example, attempting to minimize the weight of every single steel beam in a
framing system would create a large variety of different shapes, which would be difficult
to procure, transport, and install on the job site.
This pattern is sometimes applied as a substitute for the Minimum Cost pattern, but this
is not preferred. The pattern may lead to economical designs when the cost of materials
dominates over the cost of labor, transportation, and other factors. However, simply assuming that minimum weight will lead to low cost is short sighted and can lead to unintended
consequences.

Known uses
Selecting steel shapes by minimum weight is a commonly used algorithm for choosing
among the catalog of AISC shapes (Segui, 2007).

5.2.3 Minimum Cost pattern


Intent
Choose a design option that will produce the lowest overall cost.

144
Assumed background
Foundational + Mechanics + Problem recognition and solving + Design + Project management + Public policy + Business + Teamwork, and comprehension of production theory
(Koskela and Vrijhoef, 2001).

Motivation
Money is one of the most important resources on any project, and the motivation for
reducing overall cost is obvious. Yet, there are many different methods for estimating cost
and the method of arriving at minimum cost is not always clear.

Applicability
The Minimum Cost pattern can be used to choose among a set of structurally acceptable
design options that differ by their cost. The pattern is best used when the other differentiating factors in the selection are less important than cost.

Structure
The Minimum Cost pattern simply chooses the design option with the lowest cost value.

Consequences and implementation


Implementation is difficult, because there are many different methods of estimating
project and component cost. Simple cost models involve multiplying quantities by unit

145
costs. Slightly more advanced models use a unit cost function, where the unit cost is
usually high for small quantities and lower for large quantities. Mackie et al. (2008a) use
a fixed unit cost for estimating bridge repair costs, and Naeim et al. (2007) use a variable
unit cost in the PACT software for building seismic damage cost estimation.
Depending on how cost is evaluated, many different decisions might be made using
this pattern. If cost is evaluated based on up-front component costs, then the pitfall would
be choosing an option that might actually be more expensive in the long run. If cost is
evaluated based on life-cycle costs, then the initial cost might be higher, and the estimate
will also depend on risk models used.
A good approach would be to blend consideration of costs during design, construction,
structure operation, and structure functional use. The costs for design and construction
are generally an order of magnitude less than operation and use costs. The estimates for
the cost ratio between construction, operation, and business use vary from 1:5:200 (Saxon,
2005), to a perhaps more realistic figure of 1:0.4:12 (Hughes et al., 2004). Design costs on
most projects are approximately 2-5% of construction costs. So, there is a clear incentive
to consider the total project cost instead of only design fees and construction expenses in
making decisions.
Making decisions purely on cost alone often neglects other factors that are important
for the structure's performance. For example, other measures like safety are not always
easily evaluated by money costs. Speed may be more critical for structure use than cost, as
in the case of many emergency infrastructure repairs for bridges and highways.

146
Known uses
Uses of this pattern are difficult to find in most design methods, because of the difficulty
involved in computing the costs of design options. Cost tends to be evaluated after decisions
are made instead of before initial decisions. The concept of value propositions attempt to
overcome this barrier by providing tools for estimating costs earlier in the design process
(Parrish et al., 2008b).
Mackie et al. (2007) proposes a method for choosing among different bridge column
designs using total highway bridge cost including initial construction and post-earthquake
repairs.

5.2.4

Optimization pattern

Intent
Choose a design option based on some optimal value of multiple performance measures
involving tradeoffs between them.

Assumed background
Foundational + Mechanics + Problem recognition and solving + Design + Teamwork.

Motivation
Many decisions have neither a single solution, nor a solution based on one single measure of performance. For these cases where tradeoffs need to be made among multiple En-

147
gineering Performance and Social Performance measures, some method needs to be used
to make rational choices among sets of design options. The Optimization pattern describes
methods for making these choices where there may not be a single clear option, or a single
decision variable. For example, the selection of a structural system might involve several
different factors ranging from engineering criteria like drifts, accelerations to social criteria
like cost, labor rates, and material availability.

Applicability
This pattern is applicable in situations where there is no single "best" solution, but all
options involve a series of tradeoffs. This occurs most commonly the earlier phases of
design, and less so in the more detailed stages of design when most of the functional decisions have already been made. For example, the selection criteria for choosing a structural
system are not as clear cut as compared to choosing a beam size or shape.

Structure
This pattern involves the evaluation of multiple performance objectives and comparing
them to one another. This is typically accomplished in one of two ways: converting the
design problem into a computable form to be optimized using numerical tools, or using a
qualitative method of comparing the differences and advantages within a set options. The
selection of the best design option will attempt to maximize the value of each performance
objective with a higher or lower weight assigned to each objective depending on their rela-

148
tive importance.
The Minimum Weight and Minimum Cost patterns are examples of Optimization for a
single variable.

Consequences and implementation


Implementing this pattern is challenging because of the difficulty in assigning weights
to each performance objective. Assigning relative importance to each objective is a subjective process, and the results of the weighting can vary from person to person. Evaluating
multiple options requires a clear understanding of the functional design requirements so
that options are considered in an equivalent manner.
Numerical methods are more suited to fulfilling multiple Engineering Performance objectives, usually related to a mechanics model. Qualitative procedures are usually better at
comparing Social Performance objectives or a combination of both.

Known uses
At the element level, Aschheim et al. (2008) presents an optimization scheme for designing reinforcement positions for rectangular reinforced concrete beams, columns, and
walls under biaxial bending and axial load. This scheme uses a conjugate gradient search
method and defines a number of spatial variables related to the position and size of reinforcing inside the rectangular section. This method optimizes for Engineering Performance and
uses an Aggregation to represent individual rebars by considering them to be "distributed

149
uniformly, or smeared, into a single, continuous plate that is inset from the edge of the
section." Additional design is still needed to choose bar layouts and bar sizes.
At the system level, Austin et al. (1987a,b) use the

DELIGHT

system (Nye and Tits,

1986) to optimize steel frame design for three criteria: minimum volume (weight), minimum story drift, and maximum energy dissipation. A single variable, moment of inertia
/, characterized the section properties of the frame elements, and section area A for truss
elements. This method produces a well-proportioned frame system, but still requires additional design selection to choose actual steel shapes to implement the abstract design.
For qualitative comparisons, the Choosing by Advantages method (Suhr, 1999) which
assigns weights to different decision criteria and scores to each option with the advantage.
The Analytic Hierarchy Process (Saaty, 1990) uses pairwise comparisons in order to weight
and score decision criteria. The option with the highest score is considered the optimal
choice.

5.2.5 Prescription pattern


Intent
Follow an established rule for making a design decision, or confirming that a candidate
design choice is valid.

Also known as
Code-based design, "cookie cutter" procedure.

150
Assumed background
Foundational + Materials science + Mechanics + Experiments + Problem recognition
and solving + Design, and specialty knowledge for applying seismic design criteria.

Motivation
Sometimes there is only one acceptable design decision for a particular decision unit
because of regulatory constraints, or the lack of other acceptable options. These kinds of
decisions are based on an accepted method or a list of acceptable choices that are constrained by factors outside of generalized performance measures.
Prescriptions also solve problems by providing maximum or minimum ranges of quantities to be used in design. This is a good way of encapsulating general constraints based
on prior experience in industry and consensus of engineering judgment.
In the design of steel moment frames, there is a prescriptive method in FEMA 350
for the design of the moment-resisting connections. This method promotes the use of
prequalified connections with specific details, because they have been tested and proven
to achieve sufficient ductility. In this case, no non-tested connections should be used in
order to prevent fracture. The allowable design choices are limited because of the limited
amount of available laboratory testing data. These constraints are due more to external
factors than ones that are intrinsic to the design itself.

151
Applicability
This pattern addresses situations where Engineering Performance is difficult to quantify,
as in the case with stress concentrations in steel moment-resisting connections. Prescriptions are also a common method of encapsulating engineering judgment and experience
from past projects. This is useful when the solutions to design problems are generally
understood and there is a good reason for limiting the range of possible solutions.

Structure
The general use of this pattern involves following a step-by-step procedure that almost
always leads to an acceptable solution. The use could also be simply to adopt existing
design information.

Consequences and implementation


The prescription pattern fulfills its objectives by providing a simple procedure that is
easy to follow and suited for most conditions. But, this simplicity is also its greatest shortcoming. Prescriptions might not be applicable in all situations, and certain edge cases
might make the prescription invalid or, in the worst case, dangerous for design.
Many prescriptions for reinforced concrete design are invalid for high strength or light
weight concrete, and need to be specifically modified or not used at all. For example, the
common approximation for concrete elastic modulus (Ec 57,000^/^T) is not valid for
light weight concrete, in which case (Ec = w,!.'533-y/^) must be used, and even that is

152
only valid for concrete weights of 90-155 lb/ft3. Another example is the minimum steel
requirement for flexural elements, which needs to be altered to accommodate high strength
concrete with f'c > 5,000 psi.

Known uses
ACI 318, Sec. 10.5 contains rules for flexural elements that specify a maximum and
minimum amount of steel in a cross section, As^m = 3^jJ-bwd > 200bwd/fy. This minimum amount of steel is intended to encapsulate good judgment in order to prevent the
dangerous condition of sudden failure when the uncracked moment capacity is greater than
the cracked moment capacity. The designer can prevent this type of failure by simply following the prescription, instead of needing to check this condition specifically based on a
mechanics model and definition of specific engineering performance criteria.
FEMA 350 gives a prescriptive design method for prequalified steel moment-resisting
connections. The step-by-step procedures will specify all the information needed to detail
the connection for fabrication. Staying within the bounds of these prescriptions allows
the designer to complete the task, knowing that the connections will not fracture based on
confirmation by laboratory tests.

153

5.2.6

Set Exploration pattern

Intent
Investigate the consequences of multiple different options before continuing with a single option.

Also known as
Set-based design, set-based iteration.

Assumed background
Foundational + Mechanics + Problem recognition and solving + Design + Communication + Teamwork, knowledge of set-based design concepts for structural engineering
applications (Parrish et al., 2007), and comprehension of structural system behavior (Lin
and Stotesbury, 1988).

Motivation
This decision making pattern considers several design options before making a decision. Instead of choosing a single option first then evaluating it, the set exploration uses
evaluation first, then the selection of the best option.
Point-based design involves selecting a single feasible design option that satisfies Engineering Performance (5.1.5) criteria at each step in the design process and then refining
that single design (or point) while developing more details during the design process. This

154
single design is then re-worked until a solution is found that is feasible. The first design
thus selected by a structural engineer tends to be uninformed by the expertise of rebar fabricators, placers, and concrete suppliers who will perform the actual rebar detailing, rebar
placement, and concrete placement.
In contrast to point-based design, set-based design focuses on keeping the design space
as open as possible for as long as possible. A key to the success of set-based design is
knowledge sharing; whenever the feasible design space is reduced, the reason for eliminating any part of it needs to be documented and made accessible to all relevant stakeholders. Preserving the maximum number of feasible designs as long as possible reduces the
likelihood that rework will be necessary and allows consideration of Social Performance
objectives (5.1.6) by creating a process for all project participants to leverage their unique,
individual, and team-based expertise to make the project successful.

Applicability
Set Exploration can be applied in situations with low ambiguity in the decision making
process, and is usually desirable when the costs of exploring options is less than postponing early commitment, as in point-based design (Terwiesch et al., 2002). It addresses the
shortcoming of other decision making patterns by allowing several options to be carried
forward instead of having to backtrack and perform rework later in the process.
Set Exploration works well in cases where it is possible to explore different design
alternatives. When Engineering Performance criteria are unable to distinguish between

155

(18")
47cm-

iwTv
^ ^ ^ ^ ^

V u ^ T

^ ^'' *
Beam I

*8b" w ^ WPi * P v J O |
3#7bars

VlU^T

' ' * * a'


Beam 2

\ ^ ) r

w*

Beam 3

*H&T

fefe^K

' ' ' ' * * * I LJ- 1 -^

Beam 4

Beam 5

Beam 6 .

Hard Constraints:
ACI Code
Soft Constraints:
Design Preferences
Column

Column 2

Column 3

Figure 5.5: Set Exploration involves evaluating multiple options for design preferences in
addition to checking structural engineering criteria (Parrish et al., 2007).
multiple options, set exploration provides an opportunity to check downstream design criteria before making a commitment. Set Exploration does not make sense when evaluating
multiple options is too difficult because of unclear Engineering Performance criteria.

Structure
The structure of set-based design involves enumerating several options that all meet the
required Engineering Performance criteria and then evaluating them for other selection criteria. Once the options are evaluated, the optimal choice is made. An example illustrating
the selection of reinforcing steel for a beam-column joint is shown in Figure 5.5 (Parrish
et al., 2007).

156
Consequences and implementation
Exploring the set of options accomplishes a better design because it allows better consideration of downstream factors for the construction process. Using set exploration allows
better evaluation of Social Performance criteria by developing multiple options instead of
just one. This provides more information for tradeoffs and provides more information during the early design phases of a project.
Set exploration requires additional up-front work in order to develop the set of options,
and additional analysis needed to evaluate the multiple options. The implementation of setbased design is difficult because most existing design procedures depend on point-based
decision making patterns.

Known uses
Parrish et al. (2007) explains the application of set-based design strategy for structural
engineering in reinforced concrete. Set-based design is described by Ward et al. (1995)
in the context of set-based concurrent engineering. Set-based design has also been documented as a key component of Toyota's success in the automotive industry, where it is
used internally to evaluate design alternatives, and also by its suppliers (Liker et al., 1996).
The basic principles behind Toyota's practice of set-based concurrent engineering involve
mapping the design space, integrating by intersection, and establishing feasibility before
commitment (Sobek et al., 1999).
The viscous damping wall structural system for the Cathedral Hill Hospital in San Fran-

157
cisco, California was chosen based on a set-based evaluation of multiple different options
(Parrish et al., 2008a). The driving factors for choosing the system involved constructibility,
cost, seismic performance, and site conditions.

5.3 Metadata for design process and design product


The patterns described above can be used to define a metalanguage to describe the tacit
knowledge and heuristics of structural engineers, and the information about the design itself
using the definitions from Chapter 3. This language can carry data and information over
from one design step to the next, beyond what is usually transferred. This metalanguage
defines the metadata for both the design process and the design product. This complements
the language developed to store structural-state data (Stojadinovic et al., 2004; Wong and
Stojadinovic, 2004) from sensors and inspector observations. The new metalanguage and
the state data language combined allow a way to capture information about the structure
through its entire life cycle from initial concept to ongoing operation with greater depth
than previously possible.
The metalanguage consists of terms that describe the characteristics of structural design information and indicate how things are related to each other. It defines specifically
the level of intentional precision (detail) by providing a slot for the expression of the abstractions that led the engineer to design those pieces of information. When the set of abstractions are given, and the problem solving method, and choice rationalization are made

158
explicit, then the level of intentional precision is more clear.
Describing the process for arriving at a conclusion may be just as important as the
conclusion itself. Understanding why decisions are made provides the necessary context to
the decision itself. What might seem like an arbitrary choice, might be understood as the
best possible decision once the context is known. For example, a material shortage may
have driven the decision to choose a steel shape that otherwise would be considered less
desirable.
The metalanguage also reveals specific details that are not considered in this particular
selection of information. This can be done by specifying the conceptual scale and level of
grouping (agglomeration and aggregation) that the information depends upon.
The concepts behind the metalanguage are not at all new. These things have been part
of engineering design. What is new is the formalism and explicitly-named categories of
information. This formalism allows computer software to interact with design information
in ways that match the structure of how structural design really happens.
Another goal of the metalanguage should be to make it minimally burdensome for an
engineer to think in terms of it and to express knowledge either through a computer or
through a conversation. The metalanguage does not perform any thinking, but rather simply
provides slots to ensure that critical information is not inadvertently left out of the picture.
This is similar to the way template forms are. Fields in the forms indicate the information
necessary to support a process. Empty fields on the forms indicate incomplete information,
but the person filling out the form is not obligated to fill out all the information. Some

159
of it might be not applicable, or the person might simply choose not to provide all the
information. Depending on the process, fields with "N/A" inside, or blank fields might be
acceptable.
The metalanguage provides a system for capturing the tacit information that goes into
the design process. The examples that follow illustrate ways that the metalanguage can
capture decisions which must be made during the design process. In a sense, the metalanguage provides a mechanism for built-in commenting and documentation during a design
process. It allows the designer to leave "breadcrumbs" that allow him to retrace steps and
identify places where the solution can be revisited.

5.4 Examples of design and decision patterns


Using the patterns identifies the locations where cost considerations can be inserted
along with existing textbook design methods. Where design decisions are made, the additional selection criteria need to be evaluated and considered during the selection process.
Structural design requires the manipulation and thought about abstract entities, based on
mechanics theory, in order to check physical designs for suitability in their environmental
conditions (e.g. loads). These abstract entities are based on physical objects, but only on a
specialized set of their physical properties. The description needed of the physical is special
to the formation of abstractions for engineering computation and design integration, and not
al all the same as needed for construction. For example, the field iron worker would not be

160
concerned about the value of p^ or some quantity such as the ductility p.
Identifying the needs of the designer apart from the needs of physical space gives rise to
the connection between the notion of design abstractions and other related concepts. These
abstract design quantities are needed to describe the relationship between "performance"
and the physical nature of the building or structure. Theories of mechanics, analysis, and
design are only able to accomplish operations on the parameters that it employs. Detail
beyond the amount capable of handling by the theory of mechanics used is unable to be
accounted for, and really unnecessary for the theory to be employed.
Mechanics models operate at a certain minimum and maximum level of detail. The
level of detail can be chosen, and generally it is preferred to keep model at the simplest
level of detail needed to answer the design or analysis question at hand. The level of detail
needed by the model's abstractions defines the set space for design. As long as the abstract
criteria are satisfied in the model, then any detail can work properly in the design.
The model to be used depends on the decision that needs to be made. The decision unit
is a question of performance that must be answered by specifying information about the
building. The type of decision being made will affect the information that is required to
know how to answer, and the form of the answer. Some decision units will require more
information on the functional level, while some will require more information on the detail
level. It might require the resolution of ambiguities.
The problem solving method leaves room for set-based exploration, or new problem
solving methods can be developed in order to enable set-based design. Existing problem

161
solving methods based on mechanics theory already permit much freedom in set exploration.

5.4.1 Scope of examples


The scope of these examples is limited to a class of low ambiguity structural engineering design situations, corresponding to high uncertainty/low ambiguity on the uncertainty/ambiguity matrix in Schrader et al. (1993). These examples represent cases where
the variables and the functional relationships between the variables are known, but the values of the variables are unknown. High ambiguity structural engineering decision situations
are characterized by an unknown problem-solving algorithm, unknown functional relationships between the variables, or both, and are outside of the scope of these examples. These
examples are suitable for solution by either iterative problem solving methods, or set-based
strategies, because they are unambiguous and lack only maturity (Terwiesch et al., 2002;
Giletal.,2008) 1 .

5.4.2 Reinforced concrete beam design


Decision unit
What reinforcing steel is needed for a rectangular concrete beam with known loading?

'Terwiesch et al. (2002) refers to maturity as "uncertainty", a concept discussed in Chapter 3.

162
Problem solving algorithm
Reinforced concrete design using LRFD, MacGregor and Wight (2005, p. 148-155)
Lesson
The problem solving algorithm provides a method for checking a solution but not
for choosing among options. Therefore the Pick Any decision pattern is used for the
selection of bar layout.

MacGregor and Wight (2005) suggests one design procedure for beams having known
dimensions b and h. Table 5.1 presents the steps implied by this this design procedure. The
reinforced concrete beam is designed using designed using LRFD, satisfying the design
equation
<$>Mn>Mu

(5.1)

Supposing that the overall height h and width b of the rectangular beam are known, then
the design example requires the following detail and prerequisite abstractions. No other
parameters are included in the mechanics design models used for this design decision.
1. Mu: factored moment
2. d: effective depth
3. As,ASjmin: cross-sectional area of rebar
4. fy: steel yield stress
5. f'c: concrete compressive strength
6. <j>: resistance factor

Table 5.1: Design procedure for known b and h.


#

Step description

Patterns applied

1 Estimate the factored moment, Mu.

Eng. Performance (5.1.5)

Compute the effective depth, d.

Abstraction (5.1.1)

Compute the area of reinforcement, As.

Aggregation (5.1.2)

Try a specific configuration for As.

Pick Any (5.2.1)

Check whether As > A,s-,min.

Prescription (5.2.5)

Check whether the section is tension-controlled.

Eng. Performance (5.1.5)

Recompute Mn and Mn.

Updating (5.1.4),
Eng. Performance (5.1.5)

8 Recheck the area of steel required.

Updating (5.1.4),
Eng. Performance (5.1.5)

164
Here, we choose to work within a set of details needed to compute the value of Mn.
Another set of details could have been used to generate the same abstraction. Thus there is
some design flexibility to choose how open the design will be at the point where abstract
design targets are defined. Recognizing these decision implicit in the selection of theory
would avoid fixation.
Using the example in MacGregor and Wight (2005, p. 148-151) we have the following
design parameters for reinforcement selection: fy = 60,000 ksi, f'c = 3,000 ksi, b = 24 in.,
h = 24 in., and Mu = 541 kip-ft.
Step 1: Estimate factored moment demand, Mu.
The factored moment demand is estimated based on superimposed service dead loads
and live loads. In the case of known b and h, the dead load can be estimated using an
average unit weight of reinforced concrete. This estimate would need to be updated for a
design case with unknown b and h. The example uses a value of Mu = 493 kip-ft.
Step 2: Compute the effective depth, d.
The effective depth d, the distance from the top of the beam to the centroid of the
reinforcing steel, can only be estimated because the dimensions of the rebar have not yet
been chosen. The example makes a choice to design for a single layer of rebar, which
allows for a rough estimate of

drah-2.5"

= 21.5"

(5.2)

165
Step 3: Compute the area of reinforcement, As.
Moment equilibrium is used to estimate the required area of reinforcement As. For this
computation, the distance jd between the tension and compression resultants is needed. A
decision has to be made to estimate j PS 0.875. The required area of steel is then

M
A, = ^ - = 5.83in.2

(5.3)

Step 4: Try a specific configuration for As.


The example enumerates three possible choices, and uses the Pick Any pattern and
selects the second option:
1. Ten No. 7 bars
2. Six No. 9 bars
3. Four No. 9 bars + Three No. 8 bars
At this point, the mechanics model and design procedure is not sharp enough to discern
the consequences of choosing #7 bar, #8 bar, or #9 bar. Choices like this have little impact
on engineering performance and are not reflected in the model, but could have significant
impact on social performance measures. Other information such as rules of thumb, value
propositions Parrish et al. (2008b), or rebar fabricator comments are needed to make this
decision.
At this stage in the design process, before the selection of an actual rebar configuration,
the communication of preliminary information must be done in a way that does not imply

166
a level of spatial or material detail greater than what is actually present. The functional and
abstract detail is complete to the extent allowed by the problem solving algorithm.
Step 5: Check whether As > As m i n .
The selected value for As must be greater than the minimum steel area A.vmjn given by
the ACI code

V / L ^ 200bwd

As,min = -^bwd

> ^ - = 1.72 i n /

(5.4)

Step 6: Check whether section is tension controlled.


The selection of As must also satisfy the tension controlled criteria fs fy, which places
an upper limit on reinforcement. One method of checking this is to compute pt,, the reinforcement ratio at balanced failure where concrete crushes and steel yields at the same time.
This ratio can be computed using

For typical values of ecu = 0.003 and Es = 29,000ksi the expression can be simplified
with fy in units of ksi to

085p,/; /
pt =

87,000

-^ll77JooTAJ

\
<5 6)

'

This can also be expressed as the depth of the equivalent rectangular stress block in a
rectangular beam at the balanced condition

167

ab

= pl

87.000

U,ooo+/J

(5J)

Another determination of the tension controlled limit is the ratio between the depth a
and the distance dt from top of beam to extreme tension layer

= 0.3750!

(5.8)

So, with these two constraints the example gives < ^ = 0.503 and | < ?f- = 0.319.
This indicates that the section is tension-controlled for the choice of As.
Furthermore, the ACI code limits p < 0.75p/,, and a common rule of thumb says that
rebar is easier to place when p < 0.50p^.
Step 7: Compute Mn and <|>Afn.
The final rebar configuration must be checked to confirm that Mn > Mu using the
tension-controlled reduction factor value = 0.90. These constraints provide enough information for the selection of bar sizes and bar count to answer what rebar is needed for
this reinforced concrete beam.
Step 8: Check area of steel required. A final evaluation is needed to verify whether the
final design meets all the required performance criteria. Only a check on general behavior
is needed, because the minor changes in yield displacement associated with final detailing
are usually not of significance (Aschheim, 2002).

168

5.4.3 Preliminary frame design


Decision unit
What are the approximate sizes of members needed for a particular building configuration with a steel moment-resisting frame?
Problem solving algorithm
Target drift and yield mechanism (Lee and Goel, 2001).
Lesson
The frame design method uses the Updating pattern to improve initial estimates of
beam and column moment demand. This type of strategy arrives at a workable solution, but might miss a better solution. The Minimum Weight pattern is implied by
method for choosing steel shapes.

The steps in the target drift and yield mechanism method (Lee and Goel, 2001, p. 168)
are outlined in Table 5.2.

Table 5.2: Target drift and yield mechanism method.

Step description

Patterns applied

Setup

la

Estimate fundamental period

Eng. Performance (5.1.5)

lb

Estimate yield drift

Eng. Performance (5.1.5)

lc

Select target drift

Eng. Performance (5.1.5)

169

Table 5.2: (continued)

Step description

Patterns applied

Id

Calculate y

Abstraction (5.1.1)

le

Calculate p\-, A,,

Abstraction (5.1.1)

Design forces

2a

Calculate design base shear, V

Eng. Performance (5.1.5)

2b

Design forces, Fi

Abstraction (5.1.1)

Beam design

3a

Select Mpc

Abstraction (5.1.1)

3b

Calculate required beam strengths

Abstraction (5.1.1)

3c

Determine beam sizes

Minimum Weight (5.2.2)

Column design

4a

Select overstrength factors, %-t

Abstraction (5.1.1)

4b

Calculate updated forces F-m

Updating (5.1.4)

4c

Calculate design moment Mc(h) and axial forces

Eng. Performance (5.1.5)

Pcih)
4d

Determine column sizes

Minimum Weight (5.2.2)

Verify with nonlinear static and nonlinear dy-

Eng. Performance (5.1.5)

namic analyses

170

This example involves determining preliminary steel member sizes for a steel momentresisting frame in a building. The column spacing and floor heights are already established.
The next step in the design is to consider the possible sizes of steel elements that can be
used for this building. The loading is already known because of the site conditions and the
building occupancy type gives the loads. The decision is to determine approximate sizes of
a steel moment-resisting frame for a 6-story office building.
Step 1: Setup.
This step involves the use of the Engineering Performance (5.1.5) and Abstraction
(5.1.1) patterns. The engineering performance variables determined here are the building fundamental period T, yield drift 0^, and target drift 0f. The creation of two abstract
parameters y and (3/ is also used to facilitate the design thought process.
The building has five 30' bays and floor heights of 14', except for the ground floor which
is 18' high. The loads for each floor are DL = 100 psf and LL = 40 psf, and at the roof DL
= 75 psf and LL = 40 psf. The floor area is 22,500 ft2. The live load seismic contribution
a = 25%, which gives a seismic story weight w; per bay of 459 kips and 360 kips at the
roof. The structure period is estimated as T = 0.035H015 1.006 sec. The target drifts
are total drift 6, = 3%, yield drift Qy = 1%, and plastic drift Qp = 6f - 0y = 2%. Because
T > 0.7sec, then the structure ductility is estimated to be fis = 1 + -^ = 3, and R/j = /JS. The
shear proportioning modification factor is then y = ^

~ 0.556.

171
The lateral load distribution on the building is estimated using the mode shape factor b
and shear proportioning factors (3/ and A,,- for computing the base shear.

A=I'^*V
^=(Pi-p'-')|i^1

(5 io>

'

Step 2. Design forces.


The total base shear Vi, can be computed using

Vb _ ~a+yJa2 + 4yC^
W ~

(5.12)

The Ce factor depends on the building's importance factor /, seismic zone coefficient Z
and soil factor S

71S

Ce = 1 . 2 5 ^ 3

(5.13)

The total base shear for this example building is VJ, = 246kips. The base shear is distributed as Fj = XiVt,.

172
Step 3. Beam design.
Once the lateral force distribution is known, the next step in the problem solving method
is to begin designing the beams. The beams are specified by a required value of plastic
modulus Z based on story moment demands. The moment demands are computed using

MPbi=

(5.14)

where the reference beam plastic moment is

(rj=i
Mpbr = ~

Fjhj)-m
;R

(5.15)

MP = H ^ 1

( 5,6)

yn

and bottom story column demand

At this point, the set of possible steel beams is narrowed to the ones that satisfy the
moment demands. To further narrow the set, ASTM A992 grade steel is chosen. A992 is
the best grade of structural steel for seismic performance, and there are no other feasible
market alternatives. This gives a yield stress of Fy 50ksi and required plastic modulus
of
Zmin = Mpb/Fy. The beams must also meet the seismic slenderness requirements for the
flange Xpf and web Xpw.
Step 4. Column design.
Once actual beam sizes are selected, then the story moment demands Mpt are updated
using the the actual shapes Mpi, = FyZ. Then, the design procedure goes through the se-

173
lection of the columns. The first step is to select the overstrength factors ,- at each story.
For this building, the overstrength factor is , = 1.05 except at the roof where %n = 1. The
roof level force Fnu (n is for the roof level, u means updated with the selected beams) is
determined using

nu =

2Mpc +
2(ZU^jMpbj/
x^n Ta 7~a T\

(5-17)

The individual lateral story forces using the updated values are

Fiu = (ai-^+l)Fnu

(5.18)

The columns can be treated as free body cantilevers and their moments computed using
equilibrium equations

McMQ

= [ ZSkMpbA

- ^ (j^SiFiuihi-h)

(5.19)

McM(h) = 2 I bfaMpbi I " f E Wiu (hi - h) J

(5.20)

The axial forces are also computed using equilibrium and the controlling LRFD load
combination by

Pcinti = t,0.5L2

(l.2DLj + l.6LLj)

(5.21)

174

Pc,^ = 0.5PcMi +
j=i

S;

(5.22)

This provides all the loading information necessary to choose shapes for the columns
and the beams. A final check whether each joint satisfies Y,Mpc/Y,Mpt, > 1.0 ensures
strong column, weak beam behavior.
Step 5. Verify with nonlinear static and nonlinear dynamic analyses.
This final step in the design procedure verifies the final design by checking with more
detailed analysis as needed. This is intended to remind the designer to evaluate any remaining Engineering Performance (5.1.5) criteria that have not been verified in the previous
steps.

5.4.4 FEMA 350 steel moment connections


Decision unit
How should the connections be detailed in order to accomplish a moment-resisting
steel frame?
Problem solving algorithm
Design method outlined in FEMA 350Recommended Seismic Design Criteria for
New Steel Moment-Frame Buildings (SAC Joint Venture Partnership, 2000).

175

Table 5.3: Connection design procedure based on FEMA 350 basic design approach.
#

Step description

Patterns applied

1 Frame configuration

Mechanics Model (4.2.5)

Choose connection configuration

Pick Any (5.2.1)

Design connection using plastic hinge location

Eng. Performance (5.1.5),

and connection demands

Prescription (5.2.5)

Detail the connection

Prescription (5.2.5)

Lesson
Different mechanics models are needed during different stages of steel moment frame
design. A model with highly precise spatial information is needed when designing
the connections, but another model is needed to design for the overall building behavior. The method uses a combination of Pick Any and Prescription decision patterns
in order to arrive at a reliable design supported by laboratory experiments.

FEMA 350 explains a procedure for the design of new steel moment-frame buildings,
with particular emphasis on the design of the moment-resisting connections. The basic design approach is summarized in Table 5.3, and the following example is given for the design
of a reduced beam section (RBS) prequalified connection using the method summarized in
Table 5.4.

176

Undeformed
frame

thrift anqle - 0

Figure 5.6: Inelastic behavior of frames with hinges in beam span (FEMA 350, p. 3-3).
Step 1: Frame configuration.
The first step of the design procedure uses a simple Mechanics Model (4.2.5) to illustrate the design proper frame proportions. The performance criteria are primarily conceptual in this step. The mechanics model used here does not require a very high level of
maturity, and only needs basic spatial and material information in order to proceed. The
main requirements are having appropriate plastic hinge locations in the beam span (Figure 5.6) and the ability to achieve the require interstory drift angle.
Step 2: Select connection configuration.
This step in the procedure involves selecting a configuration that will be compatible
with the structural system, fit the element sizes, and be able to provide adequate inter-

177
story drift capacity. Some connections are only permissible for certain ranges of beam
and column size because of limited experimental data. If these criteria are met, then any
prequalified connection is considered acceptable, and the Pick Any decision pattern (5.2.1)
can be applied to choose among the catalog of prequalified connections. This example uses
the reduced beam section (RBS) prequalified connection.
Step 3: Design connection using plastic hinge location and connection demands.
The plastic hinge location s/, (Figure 5.7) is spatial design information based on the
selection of connection configuration from Step 2. This Engineering Performance parameter also depends on the the amount of gravity load placed on the beam. When the gravity
load is low, the maximum moment is expected to occur at the column faces; when high,
the maximum moment could occur somewhere inside the beam span, requiring additional
computation. After sj, is determined, the other Engineering Performance (5.1.5) parameters are computed: the probable plastic moment at hinges Mpr, shear at the plastic hinge
Vp, moment strength demands at the column face Mf and column centerline Mc, and the
yield moment at column face Myf.
Specifically for RBS design, the location 57, determines the position of the flange reduction. Then the rest of the prescriptive process outlined in Table 5.4 determines the
additional spatial information for connection details (Figure 5.8).

178

Connection
reinforcement
(if applicable)

Educed beam
section
( i f applicable)

Figure 5.7: Location of plastic hinge formation (FEMA 350, p. 3-6).


Table 5.4: Design procedure for Reduced Beam Section (RBS) steel moment-resisting
connection.
#

Step description

Patterns applied

1 Determine length and location of flange reduction

Prescription (5.2.5)

Determine depth of flange reduction, c

Prescription (5.2.5)

Compute Mf and Mc

Eng. Performance (5.1.5)

Calculate shear at column face, Vf

Eng. Performance (5.1.5)

Design shear connection of beam to column

Prescription (5.2.5)

Design panel zone

Prescription (5.2.5)

Check continuity plate requirements

Prescription (5.2.5)

Detail the connection

Prescription (5.2.5)

179

4c2+ b2

Figure 5.8: Reduced beam section connection detail (FEMA 350, p. 3-40).

180

5.5

Summary

What I have done in this chapter:

1. Documented existing methods for design analysis and decision making as patterns.
2. Defined a vocabulary of pattern names for describing structural engineering design
methods, which can be used as a metalanguage for better describing design intent.
3. Illustrated pattern usage for steel and reinforced concrete design at the system, element, and detail scales.
4. Indicated places in design procedures where decision making can be improved.
5. Explained the reasons for using different decision patterns based on available information and the information type to be determined.

181

Chapter 6
Extending performance-based design
Performance-based design has proven difficult to transfer into practice because contemporary design procedures are not suitable for supporting all aspects of a complete
performance-based methodology. Many design procedures have been proposed for implementing performance-based methodology, but typically suffer from common flaws: a
narrow definition of performance that focuses only on the as-built condition but neglects
the other project phases, lack of guidance on how to select among multiple design options,
and an emphasis on design evaluation without a process for design creation. This chapter
extends performance-based design processes by using the patterns that solve these common
flaws.
Existing design methods can be augmented in order to consider performance criteria
for all phases, provide more guidance on design selection, and provide a method for evaluating and generating design options. These extensions to the performance-based design

182
process can be accomplished using the appropriate design patterns within structural engineering design procedures. This chapter investigates and provides examples of three design
procedure extensions: (1) additional analysis for decision support, (2) changing decision
patterns, and (3) using set-based decision patterns.

6.1

Extension 1: Additional analysis for decision support

Problem
Performance is not considered for all project phases.
Solution
Add additional analysis that considers performance in the lacking phases.
Description
Adding additional analysis for decision support improves design by (1) providing performance criteria for all the project phases instead of only considering the as-built structure,
and (2) providing additional information to discern benefits among multiple options (Figure 6.1). Many current design procedures only consider Engineering Performance (5.1.5),
which are insufficient to differentiate between design options that may have very different
Social Performance (5.1.6). These additional analysis steps may require additional work,
but ultimately allow better decisions in the design process. Adding additional analysis
might require a change in decision pattern in order to maximize or minimize a different
objective, or to attempt optimizing among a set of different tradeoff criteria.

183

Engineering Perf.
criteria

Engineering Perf.
criteria

Social Perf.
criteria

[Decision pattern]

[Decision pattern]

Acceptable
final design

Better
final design

Figure 6.1: Additional analysis provides more performance criteria for making a better
final design choice.
Two examples illustrate this extension: (1) including construction phase performance
in reinforced concrete beam design, and (2) including construction and repair phase performance in reinforced concrete bridge column design.

Example 1: Labor rates in reinforced concrete elements


Additional analysis is inserted into the reinforced concrete beam reinforcement procedure described in Sec. 5.4.2. The original design procedure did not consider any social
performance design criteria and only focused on engineering performance based on flexural
resistance. There was no additional information that could be used to discern the benefits
of the three different options considered for bar layout. The improvement adds a Social
Performance requirement, labor productivity rate, to refine the selection of the bar layout
(Figure 6.2).

184
The effect of this additional analysis will sharpen the selection of bar layouts. Supposing that a minimum labor productivity rate of 2,000 lbs/man-day were required for the
beam rebar, only the options using #8 and #9 bar would be feasible. The labor productivity
of straight #7 bars is on the order of 1,700 lbs/man-day, while #8 and #9 bar are higher than
2,000 lbs/man-day (Bennion, 2007; Parrish et al., 2008b). The labor placing rates for shear
walls and beams are shown in Figure 6.3. This eliminates the feasibility of using ten #7
bars.
This additional analysis will ensure a selection that satisfies performance in the construction phase of the project in addition to the engineering performance of the as-built
structure. The same Pick Any decision pattern can still be used to make the final decision
on rebar layout.

Example 2: Bridge repair costs and repair time


The existing method for designing reinforced concrete columns for highway bridges
emphasizes performance in the as-built condition. Adding additional analysis for evaluating bridge repair cost and repair time provide better performance-based design by considering performance during the bridge's construction and repair phases. This can be accomplished by integrating the LLRCAT methodology (Sec. 4.3) and rebar labor productivity
rates into the design procedure.
Some of the steps involved in the column design for a single column bent, 2-span box
girder bridge are summarized in Table 6.1. This procedure is based on information from

185

Existing method

Improved method

3. Compute required
reinforcement area, As

3. Compute required
reinforcement area, As

3a. Set a minimum


required labor
rate for installation

4. Try a specific
configuration for As

<= Additional analysis

4. Try a specific
configuration for As

Figure 6.2: Additional analysis considers construction phase performance by analyzing


labor productivity rates during reinforcement design.

186

1,800

1,600

1,400

1,200 4 - C

1,000

500 (.78 in2)

1000 (1.55 in2) 1500 (2.33 in2)

2000(3.10 in2)

2500(3.88 in2)

3000 (4.65 in2)

Figure 6.3: Labor productivity rates for shear walls and beams vary by bar size and bending
type (Parrish et al., 2008b).

187

Table 6.1: Original procedure for single column bent design.


#

Step description

1 Determine column size limits based on super-

Patterns applied
Prescription (5.2.5)

structure design
2

Choose column size

Pick Any (5.2.1)

Determine column reinforcement requirements

Prescription (5.2.5),
Eng. Performance (5.1.5)

Choose longitudinal bar layout

Pick Any (5.2.1)

Determine confinement requirements

Eng. Performance (5.1.5)

Choose transverse bar layout

Pick Any (5.2.1)

the column design procedure outlined in Caltrans Bridge Design Aids (BDA), Ch. 15 and
Ketchum et al. (2004), using Caltrans Seismic Design Criteria (SDC).
Step 1 determines column size limits based on the ratio between column cross section and bent cap depth in order to meet joint shear requirements. This limits acceptable
columns to the ratio 0.7 < ^ < 1.0 (Caltrans SDC, Eq. 7.24). This provides basic prescriptive criteria, but provides no additional guidance beyond this point. Performance for
the repair phase of the structure's life can be considered by limiting the maximum repair
cost ratio for the structure. Knowing the superstructure design and the basic estimates on
column size are sufficient to employ the LLRCAT methodology for estimating the cost of
repairing the bridge after a major earthquake.

188
Steps 3 and 5 establish requirements for longitudinal reinforcement and transverse reinforcement requirements. The required longitudinal reinforcement ratio can be determined
from design charts in Caltrans BDA, and the transverse reinforcement spacing and bar size
determined based required shear strength in the plastic load case. Beyond these minimum
limits, the procedure gives no additional guidance on rebar combinations that are the best
for construction phase performance. The construction phase performance can be considered by setting a limit on the minimum labor productivity rates for different types of rebar
type, size, and layout as in the previous beam example.
The improved procedure with these additional analysis steps is summarized in Table 6.2. The extra analysis steps before decisions add performance consideration for the
construction and repair phases.

6.2 Extension 2: Change decision patterns


Problem
Design methods are unable to differentiate among several choices.
Solution
Change the decision pattern to one that optimizes performance. Add analysis if performance measures are unavailable.
Description
A decision making pattern can be replaced with a better decision making pattern that

Table 6.2: Improved procedure for single column bent design.


#

Step description

Patterns applied

Determine column acceptable column size range

Prescription (5.2.5)

based on superstructure design


la

Limit column size range by maximum accept-

Eng. Performance (5.1.5)

able repair cost ratio size range


2

Choose column size

Pick Any (5.2.1)

Determine column reinforcement requirements

Prescription (5.2.5),
Eng. Performance (5.1.5)

3a

Limit reinforcement options by minimum re-

Social Performance (5.1.6)

quired labor productivity rate


4

Choose longitudinal bar layout

Pick Any (5.2.1)

Determine confinement requirements

Eng. Performance (5.1.5)

5a

Limit reinforcement options by minimum re-

Social Performance (5.1.6)

quired labor productivity rate


6

Choose transverse bar layout

Pick Any (5.2.1)

190
targets the desired measures of performance. These replacement decision patterns often
take advantage of the information from additional analysis (Figure 6.4).
When substituting the Pick Any pattern for another decision pattern, it is best to combine with additional analysis that will enable the support of a different decision pattern.
Usually the additional analysis will be for a Social Performance analysis because these are
the types of performance criteria that are not usually inside most design procedures.

Example: Choosing steel connections


The example of selecting steel moment connection types in the example from p. 174
is improved by changing the decision pattern and adding additional analysis for the cost
of each connection. The design method suggested by FEMA 350 provides no guidance
on which moment-resisting connections should be chosen, beyond the limits on acceptable
engineering performance. The design method is unable to provide additional information
to choose, and therefore the Pick Any pattern could be used.
A better alternative would be to use the Minimize Cost pattern and perform additional
analysis to investigate the costs for each valid prequalified connection option (Figure 6.4).
This could be done by considering factors for cost-effective steel building design. Carter
et al. (2000) describe several criteria for bolted and welded connections, as well as general suggestion for beam and column selection, material costs, fabrication labor costs, and
erection labor costs.
Market-specific information could be obtained from specialty connection suppliers as

191

Prequalified
connections
.

Pick Any

Evaluate cost
of connections

Minimum Cost

Figure 6.4: Replacing a Pick Any pattern with another decision pattern produces better
results and may require additional analysis.
well. For example, data from the supplier of the SlottedWebproprietary connection
shows that for an office tower with 750 moment connections cost over $340,000 less on
average than two other options for moment connections (Seismic Structural Design Associates, 2005). Using this additional analysis to choose a minimum cost option before
proceeding with design requires additional work in design, but leads to better performance
during the construction phase.

192

6.3 Extension 3: Use set-based decision patterns


Problem
An evaluation method is available to check performance, but cannot be used adequately
until more details are chosen. The structure design is not mature enough to fully estimate
performance measures, yet it is also too early to commit to a particular design choice.
Solution
Use the Explore Set (5.2.6) decision pattern to evaluate performance of multiple candidate design options, then optimize among the options to make the final design choice.
These sets of design options can be further evaluated for other measures of performance,
allowing the exploration of more options, which would have otherwise been neglected (Figure 6.5).

Example: Beam-column joint reinforcing steel selection


The selection of beam-column joint reinforcing steel illustrates the use of the set-based
design concept (Parrish et al., 2007). Constructibility and structural performance are the
main performance factors for joint design.
Mechanics models analyze the loads and provide the minimum required steel area As
for the beams and the columns (Figure 6.6). Prescriptions from ACI 315 provide additional
limits on dimensions and rebar layout. These two factors take care of the requirements for
engineering performance on the joint. But, at this stage there is insufficient information
to characterize the constructibility of the joint. The mechanics model and prescriptions

Existing method

Improved method

Engineering Perf.
criteria

Engineering Perf.
criteria

[Decision pattern]

Explore Set

Acceptable
final design

Option 1

Option

Option n

Evaluate more
Eng. and Social
Perf. criteria

Optimization

Better
final design
Figure 6.5: Substituting a set-based decision patterns allows further evaluation of mu
options in order to find a design with better performance for all project phases.

194
are insufficient for selecting a rebar configuration that considers construction performance.
Additional preferences, rules of thumb, and feedback from rebar fabricators and placers are
needed to proceed with the best option.
But, the design lacks the information needed to evaluate construction performance. The
information is not yet available because the design is not detailed enough, using aggregations of bar instead of actual bar layouts. Using the Pick Any decision pattern to choose any
layout and go proceed with design might work, but would lead to suboptimal construction
performance.
Instead, the set-based approach can be used to evaluate construction performance. Detailing requirements can be determined by setting a ductility demand and then determining
strength (Aschheim, 2002), or by setting a deformation mechanism (Lee and Goel, 2001).
Several options can be chosen from the limits of acceptable strength and ductility pairs and
other prescriptive design limits (Figure 6.7).
Then, the combinations of beam and column reinforcement can be evaluated for construction performance using preferences from the design team. The double layer reinforcement option might be eliminated due to increased labor required to place the bars. Certain
sizes of rebar might be eliminated due to material availability or other market constraints.
The exact spacing of the bars inside the joint can be considered to check whether there
is enough space to avoid congestion. This would be difficult to check without evaluating
a specific layout of reinforcing bar including bar size, bar count, and each bar's position.
After evaluation of construction performance, the best option can be chosen. Since all

195

(18")
47cm

As>l9cm2

61cm
(24")
A's>l2cm2

Seam B
(24")
61cm

11

As>6lcm2

61cm
(24")

Column C

??fe

7M7s

Figure 6.6: Required steel area based on mechanics model (Parrish et al., 2007).
the options were already satisfactory for as-built structural performance, the choice can be
made on construction performance alone.

(24")
61cm

I
1

61 cm I6#7bars
(24")

Column 2

8# 10 bars

Column C

I2#8bars

Column I

Column 3

(18")
47cmAs>!9 cm2

I
1

61cm
(24")
A's > 12 cm2

beam b

(18")
47cm-

5#7 bars

4#8 bars

3#9 bars

3#7 bars

3#7 bars

3#7 bars

6lcm
(24")

Beam I

Beam 2

3# 10 bars

7# 10 bars

Beam 3

4#8 bars
3#7 bars

Beam 4

3#7 bars

Beam 5

3#7 bars

Beam 6

Figure 6.7: Bar layout options for beam and column.

197

Summary

What I have done in this chapter:

1. Explained three methods for including performance-based criteria for all project phases into existing design procedures.
2. Provided examples of these three improvement patterns.
3. Demonstrated the applicability of using design and decision patterns to identify general cases for improvement.
4. Developed examples for analyzing design methods and improving their behavior.

198

Chapter 7
Conclusion
7.1 What I have done here
In this thesis I have defined terms describing structural engineering design information,
explained patterns of communicating design information, described patterns of structural
engineering analysis, and patterns for making decisions within design structural engineering design procedures. Several examples from current design methods illustrate the use
of these patterns. I have also suggested general recommendations for extending design
methods in order to correct common problems.
Chapter 2 explained the phases in the structure life cycle. A fully performance-centric
design methodology should address performance in all of these phases.
Chapter 3 defined different aspects of the content and quality of structural design information. These definitions were applied to explain how different types of content and

199
quality influence structural analysis, and articulated several challenges to communicating
information quality.
Chapter 4 documented patterns for communicating design information at the appropriate levels of detail. The examples illustrated the necessity of annotation and text in order to
describe structure attributes. Text and graphics combined excel at communicating conceptual physical design information. An example of performance criteria for the repair phases
was given for reinforced concrete bridges. Some comments on the use of BIM technology
indicate requirements necessary for implementing performance-based design.
Chapter 5 documented patterns for analysis and decisions. These patterns define a
vocabulary for describing structural engineering design methods, which can be used as a
metalanguage for better describing design intent. Pattern usage was illustrated for steel and
reinforced concrete design at the system, element, and detail scales. The use of patterns
helps identify the places in design procedures where decision making can be improved, and
explains the reasons for using different decision patterns based on available information.
Chapter 6 explained three methods for extending performance-based criteria in order
to avoid the common flaws of neglecting performance in some project phases, lack of
guidance for making decisions, and emphasis on evaluation alone. These methods identify
the reasons why performance-based design is difficult to implement and propose feasible
solutions.

200

7.2 Future work


The work in this thesis is never complete because there are many additional directions
to consider. A few suggestions for future work are described below:

7.2.1 Expanded pattern catalog


Forming a more comprehensive and detailed catalog of additional analysis patterns
would be helpful for understanding and teaching about the vast number of different approaches are in the literature. Specific methods for earthquake engineering analysis such
as the Yield Point Spectrum method (Aschheim, 2002) and the Performance-based Plastic Design method (Lee and Goel, 2001) could be formulated as design patterns in order
to succinctly communicate when such methods are applicable and advantageous within a
design procedure. The six methods for estimating inelastic displacement compared in Miranda and Ruiz-Garcia (2001) could also be written as patterns in order to list the specific
details involved in their implementation and when they are best applicable to use within
design procedures.
Future work could also develop additional communication patterns for the specifically
handling probabilistic uncertainty throughout the design process. Instead of handling variables in their single deterministic values, their probability distributions would need to be
understood. This would reflect the uncertainty between the real and ideal as-built states for
a fully mature design.
More patterns for extending performance-based design can help improve the methods

201
of today. Three examples were provided in Chapter 6, and many more could be developed
by analyzing where analysis is used for making decisions, and what criteria are used to
make those decisions.

7.2.2

Communication patterns with BIM

Future work could be done to extend the discussion of communication patterns into how
BIM technologies are influencing the way information is exchanged and expressed in design practice. There is great potential in using the recent advances of computer technology
for enhancing communication. But, much of the emphasis has been on the communication
of detailed spatial and material information, and not so much on developing these similar
tools for modeling at the preliminary phases of design. More work could be done to investigate how to incorporate these immature phases of design into a more unified set of
tools.
The use of BIM could also focus on modeling for the operation and repair phases of
a structure. There is great potential for merging building health monitoring data into a
cohesive framework with the building model. Instead of building information modeling,
during the operation phase, additional data could be merged with the building model to do
building information management as well. Building information management processes
might involve tracking the state of the structure after initial construction is over and the
building is placed into normal service. The effects of design and construction decisions
could be observed with the building in service. The merger of these pieces of information

202

could provide research engineers and practicing engineers with valuable data for improving
design.

7.2.3 Teaching and training engineers


Future work could explore using communication, analysis, and decision patterns as a
teaching tool that can help explain the reasons behind the steps involved in design procedures. This would benefit students by providing them with more information about the use
of each pattern so that they can learn to practice good judgment for evaluating design and
communication procedures. Future work could expand this type of analysis of communication, analysis, and decision patterns into a more complete catalog for teaching students.
A masters-level class on structural systems might include a discussion of patterns for
both the structural systems (e.g., moment frames) and the methods for designing them (e.g.,
Lee and Goel, 2001). This would provide students with clear distinctions between the benefits and challenges of using and designing each system. The pattern method could also
equip students with tools for choosing different analysis procedures available to them. For
example, there are many different numerical methods for performing nonlinear structural
analysis for earthquakes. But, many students tend to be confused about when and where
to use each one. Clear patterns that describe the strengths and weaknesses of each method
would make this topic much more clear, and enable students to make better informed decisions in design practice. The pattern template used in the classroom could later become a
handy quick reference sheet in the design office. A blank pattern template could be used on

203

an exam for testing students' conceptual understanding of structural systems and methods
for analysis and design.
An undergraduate-level class on design communication might include specific topics
involving when and why to use different communication patterns. These topics could connect to a capstone design course in order to give students an opportunity to communicate
a "real" design. For example, the project could begin with writing a conceptual narrative,
and making some preliminary sketches with annotation. Then, the level of detail would
increase throughout the course prompting the use of other communication patterns. Teaching the patterns might also give students the tools to choose methods of communication,
instead of merely learning how to implement communication, such as merely learning how
to use CAD or BIM software without also learning why and when to use it. Simply requiring students to turn in a "design report" with a list of required pieces would not be as rich
as providing the rationale for each piece using a pattern description.
The difference between teaching and training is clarified when the different patterns
are used to explain design methods. Teaching involves imparting knowledge and training
involved imparting skills (Gregory, 1886; Shafer, 1985). Thus, teaching can involve the
understanding and use of the patterns, while training can focus on the actual implementation and execution of these patterns. This distinction between the knowledge about design
and the actual work of design can benefit the way students understand and discuss what is
taught.
When van den Broek (1940) wrote on the use of plastic analysis, he noted that, "In

204

the training of engineers, the emphasis on the theory of elasticity leads to a respect for
the elastic limit which sometimes borders on fear." Spelling out the specific reasons why
certain analysis and decision patterns are used in a design procedure could help eliminate
this kind of unfounded fear that happens when changes are made. The use of the pattern
definitions and explanation of applicability and consequences can make these tacit fears
and assumptions more clear and considered.

7.2.4 Design with high ambiguity and unclear mechanics


The examples developed in this illustrate the process of set-based design under conditions where the mechanics assumptions are well-known and there are an adequate number
of abstractions that can express the performance of generic structural configuration. One
area of future work is determining how to do set-based design in situations where the performance of the system cannot be evaluated using established design methods.
The lack of ability to evaluate occurs commonly when using innovative designs. In
these cases, often laboratory tests are required to verify the performance of the new type of
structural device. It is in these situations that the tests may verify a particular instance of the
device, but not be able to verify an entire class of similar devices. The advantage of having
a well-established mechanics theory is that it allows one to make categorical assumptions
about behavior over an entire set of options even though the individual instantiations are
rare or have never been built before.

205

7.2.5 Reliability assessment during design


Future work could be done on describing patterns for reliability assessment during design. The probabilistic uncertainty of a future building's performance is categorized differently than that of an existing building. Der Kiureghian and Ditlevsen (2008) comment that
"the different categorization of uncertainties in an existing versus a future building dictates
a fundamental difference in the methods used for assessing their reliabilities." An existing building's reliability should be assessed using the known history of the building, but a
future building is known only from the design process and can be thought of as a random
sample taken from a population. For example, the concrete strength of a future building is
unknown and this aleatory uncertainty cannot be reduced. But, in an existing building, the
uncertainty in concrete strength is epistemic and could be reduced by obtaining data from
material samples obtained during construction.
Understanding this distinction leads to different patterns of analysis that are suitable
for existing or future structures. For example, a pattern could be described for Bayesian
techniques that are well-suited for information updating on existing structures, and another
pattern could account for the aleatory uncertainty between a design and the actual as-built
condition. Communication patterns could be explored for representing this distinction in
sources of uncertainty, which "has been missed in much of the literature on structural reliability" (Der Kiureghian and Ditlevsen, 2008).

206

7.2.6 Development of computer tools for supporting set-based design


Computer tools can help support set-based design by making it easy to consider a range
of options that fulfill defined performance criteria. The framework for this comes out of
identifying appropriate abstract quantities and knowing how to select a range of instantiations that satisfy the abstract constraint. This would be like making a tool to conduct an
easy "parameter" study.
For example, p is an abstract quantity derived from reinforced concrete mechanics theory, and it is manifested in reality from a particular bar configuration inside a particular
concrete element. So, a tool that would allow iteration over many real rebar configurations
and values of As for a given b and d of a beam would be helpful in evaluating options.
This concept is important because the mechanics theory does not really provide any
guidance beyond the computation of parameters like p, As and checking whether a particular real configuration satisfies the required demand and capacity requirements. Mechanics
theory doesn't provide information about the constraints on bar availability, ease of placing,
or fabricator efficiency. Allowing mechanics theory to function on the level of mechanics
theory, and allowing other factors to influence the other design decisions will allow better
designs for overall project performance. Avoiding confusion between the categories and
types of design information can be helped with a computer tool.
I have already developed a prototype tool written in C# that integrates with the Tekla
Structures OpenAPI. This tool is able to evaluate Social Performance design criteria such
as: automatic machinability, rebar placing time, and to compute information on the level of

207
1 * ki-b.H S f t I I C I V M I D I ' M U I I
Rebar

Labor

!ft' -

Inul

_ |n|x|

.'38fe

Model

J-J

button 1

Refresh

Bar size d ashboard


Materia! A v a i l a b l y

-*

10

>=n< i i

Bending tolerance

10

-n<*!*-

Cost

10

--

Placing Time

"

10

i'

X--1-' l.fffsfe
#

Gel Concrete Properties


BE1
Dimensions
J127D*312 8
Concrete strenj-Si
|4G:C

Hoop size
J2-35bar
Vertical steel ^ze
|3C =11 bar
Short dr ties
Long * . ties
|l-=Eoar

1 dSaedPrqpeifiss

3nine tools

Show Selected Obi Oass j

Wal

eetbarproperties

BE2

Dimenwn*
J9144-508
Concrete strength
(4000

Dimensions
j 1270*812 S
Concrete amngth

Honzortelbarst-e
J5S-S5bar
Vertical bar size
jSC-Pbbar

Hoop size
J2-S5bar
Vertical steel size
J30-Pl1bar

8arsize

1"

Grabbing Tekla. Structures. Model SingleRebar


Moving to next object...
Grabbing Tekte.Stajctures.Model.SingteRebar
Moving to neat object...
Grabbing Tdda.Structures.Modd.SngteRebar
Moving to next object...
Bar Size

I Bar Area j No. Bars 1 Total Seel Area

A':
!

Short dr. ties


]3-s5bar
Longdir ftes
)1-S5bar
Tsa cotorj by userfieki 3
Check machine bending

Reset color

Figure 7.1: Set-based design tools can integrate with existing BIM software. This prototype
tool uses data from a reinforced concrete shear wall model in Tekla Structures 14.
detail necessary for rebar fabrication time and cost estimates. The tool is able to extract design information from a Tekla Structures model for computation, and input the information
back into the model for visualization (Figure 7.1).

208

Bibliography
ACI Committee 315 (1994). ACI Detailing Manual. American Concrete Institute, Detroit.
ACI Committee 318 (2005). Building Code Requirements for Structural Concrete and
Commentary. American Concrete Institute.
Al Sarraj, Z. M. (1990). Formal development of line-of-balance technique. Journal of
Construction Engineering and Management, 116(4): 689-704.
Alexander, C , Ishikawa, S., Silverstein, M., Jacobson, M., Fiksdahl-King, I., and Angel, S.
(1977). A pattern language : towns, buildings, construction. Oxford University Press,
New York, xliv, 1171 pp.
American Institute of Steel Construction (2005). Steel construction manual. American
Institute of Steel Construction, Chicago, thirteenth edition.
American Society of Civil Engineers (2000). Quality in the constructed project: a guide
for owners, designers, and constructors. American Society of Civil Engineers, Reston,
VA.

209
American Society of Civil Engineers (2008). Civil Engineering Body of Knowledge for
the 21st Century : Preparing the Civil Engineer for the Future. American Society of
Civil Engineers, Reston, VA, second edition, http://www.asce.org/professional/
educ/bok2.cfm.
Anwar, N., Kanok-Nukulchai, W., and Batanov, D. N. (2005). Component-based, information oriented structural engineering applications. Journal of Computing in Civil Engineering, 16(1):4557.
Applied Technology Council (2008). Quantification of building seismic performance factors : ATC-63 project report, 90% draft. FEMA Report P695, Federal Emergency Management Agency.
Aschheim, M. (2002). Seismic design based on the yield displacement. Earthquake Spectra, 18(4):581-600.
Aschheim, M., Hernandez-Montes, E., and Gil-Martin, L. M. (2008). Design of optimally
reinforced RC beam, column, and wall sections. Journal of Structural Engineering,
134(2):231-239.
Austin, M. A., Pister, K. S., and Mahin, S. A. (1987a). Probabilistic design of earthquakeresistant structures. Journal of Structural Engineering, 113(8): 1642-1659.
Austin, M. A., Pister, K. S., and Mahin, S. A. (1987b). Probabilistic design of moment-

210
resistant frames under seismic loading. Journal of Structural Engineering, 113(8): 16601677.
Ballard, G. (2000). The Last Planner System of Production Control. PhD thesis, The
University of Birmingham, Birmingham, United Kingdom.
Ballard, G. (2001). Lean Project Delivery System. White Paper 8, Revision 1, Lean
Construction Institute, h t t p : //www.leanconstruction.org/pdf/WP8-LPDS.pdf.
Bay Bridge Public Information Office (2008).
sion span.

Online,

Fact sheet : Self-anchoring suspen-

http://baybridgeinfo.org/sites/default/files/pdfs/

sasfactsheet.pdf.
Bennion, H. (2007). Rebar placing rates, personal communication. PCS Steel, Inc., Fairfield, CA.
Birley, D. (2008). Reinforcement placing drawings are not shop drawings : requirements
for early submittal can be detrimental. Concrete International, 30(12):48-50.
Browning, T. R. (2001). Applying the Design Structure Matrix to system decomposition
and integration problems: A review and new directions. IEEE Transactions on Engineering Management, 48(3):292-306.
California Department of Transportation (2004). Seismic design criteria, version 1.3.
Engineering services manuals, Division of Engineering Services, Sacramento, CA.

211
http://www.dot.ca.gov/hq/esc/techpubs/manual/othermanual/
other- engin-manual/ seismic- design- c r i t e r i a / sole. html.
California Department of Transportation (2005). Bridge design aids. Engineering services
manuals, Division of Engineering Services, Sacramento, CA.
http://www.dot.ca.gov/hq/esc/techpubs/manual/bridgemanuals/
bridge-design-aids/bda.html.
Carter, C. J., Murray, T. M., and Thornton, W. A. (2000). Cost-effective steel building
design. Progress in Structural Engineering and Materials, 2:16-25.
Chopra, A. K. (2007). Dynamics of structures : theory and applications to earthquake
engineering. Pearson-Prentice Hall, third edition, xxxiv, 876 pp.
Christy, C. T. (2006). Engineering with the spreadsheet: structural engineering templates
using Excel. ASCE Press, Reston, VA. xv, 317 pp.
Codd, E. F. (1990). The relational model for database management: version 2. AddisonWesley. xxii, 538 pp.
Concrete Reinforcing Steel Institute (2000). Reinforcing Bar Detailing. Concrete Reinforcing Steel Institute, Schaumberg, IL, fourth edition.
Concrete Reinforcing Steel Institute (2001). Manual of Standard Practice. Concrete Reinforcing Steel Institute, Chicago, IL, 27th edition.

212
Concrete Reinforcing Steel Institute (2002). Placing drawingsthe detail drawings for
reinforcing bars in site-cast reinforced concrete construction. Engineering Data Report
EDR50, Concrete Reinforcing Steel Institute.
Concrete Reinforcing Steel Institute (2008). CRSI Design Handbook 2008. Concrete Reinforcing Steel Institute, Schaumberg, IL, tenth edition.
Cornell, C. A. and Krawinkler, H. (2000). Progress and challenges in seismic performance
assessment. PEER Center News, 3(2): 1-2.
Day, A. (1997). Digital Building. Laxton's, Oxford, xv, 176 pp.
Der Kiureghian, A. and Ditlevsen, O. (2008). Aleatory or epistemic? Does it matter?
Structural Safety, 31 (2): 105-112.
Ditlevsen, O. (1982). Model uncertainty in structural reliability. Structural Safety, 1(1):7386.
Ditlevsen, O. and Madsen, H. O. (1996). Structural reliability methods. Wiley, Chichester;
New York, xi, 372 pp.
Doumont, J.-L. (2002a). The three laws of professional communication. IEEE Transactions
on Professional Communication, 45(4):291-296.
Doumont, J.-L. (2002b). Verbal versus visual : A word is worth a thousand pictures, too.
Technical Communication, 49(2):219-224.

213
Eastman, C , Teicholz, P., Sacks, R., and Liston, K. (2008). BIM handbook : a guide to
building information modeling for owners, managers, designers, engineers, and contractors. Wiley, Hoboken, NJ. xiv, 490 pp.
Fee, G. D. and Stuart, D. (2003). How to read the Bible for all its worth. Zondervan, Grand
Rapids, MI, third edition. 287 pp.
Ferguson, E. S. (1992). Engineering and the mind's eye. MIT Press, Cambridge, MA. xiv,
241 pp.
Gamma, E., Helm, R., Johnson, R., and Vlissides, J. (1995). Design patterns : elements of
reusable object-oriented software. Addison-Wesley, Reading, MA. xv, 395 pp.
Gil, N., Beckman, S., and Tommelein, I. D. (2008). Upstream problem solving under
uncertainty and ambiguity: evidence from airport expansion projects. IEEE Transactions
on Engineering Management, 55(3):508-522.
Gregory, J. M. (1886). The seven laws of teaching. Congregational Sunday School and
Publishing Society, Boston. 144 pp.
Harris, H. G. and Sabnis, G. M. (1999). Structural modeling and experimental techniques.
CRC Press, Boca Raton, FL, second edition. 789 pp.
Heng, B. and Mackie, R. (2008). Using design patterns in object-oriented finite element
programming. Computers and Structures, in press: 10 pp.

214
Hughes, W., Ancell, D., Gruneberg, S., and Hirst, L. (2004). Exposing the myth of the
1:5:200 ratio relating initial cost, maintenance and staffing costs of office buildings.
In Khosrowshahi, R, editor, Proc. 20th Annual ARCOM Conference, volume 1, pages
373-381, Heriot-Watt University, Edinburgh, Scotland. Association of Researchers in
Construction Management.
Imhof, E. and Steward, H. J. (1982). Cartographic relief presentation. De Gruyter, Berlin;
New York, xviii, 389 pp.
Jansson, D. G. and Smith, S. M. (1991). Design fixation. Design Studies, 12(1):3-11.
Kang, G. S. and Kren, A. (2006). Structural engineering strategies towards sustainable design. In Proc. Structural Engineers Association of Northern California 2006 Convention,
pages 473^190.
Kardon, J. B. (2003). The standard of care of structural engineers. PhD thesis, University
of California, Berkeley.
Ketchum, M., Chang, V., and Shantz, T. (2004). Influence of design ground motion level on
highway bridge costs. Lifelines Project 6D01, Pacific Earthquake Engineering Research
Center, Berkeley, CA.
Kneer, E. and Maclise, L. (2008). Consideration of building performance in sustainable design : a structural engineer's role. In Proc. Structural Engineers Association of Northern
California 2008 Convention. 15 pp.

215
Koskela, L. and Vrijhoef, R. (2001). Is the current theory of construction a hindrance to
innovation? Building Research and Information, 29:197-207.
Lee, S.-S. and Goel, S. C. (2001). Performance-based design of steel moment frames using
target drift and yield mechanism. Research Report UMCEE 01-17, Department of Civil
and Environmental Engineering, Univeristy of Michigan.
Liker, J. K., Sobek, II, D. K., Ward, A., and Cristiano, J. J. (1996). Involving suppliers in
product development in the United States and Japan : evidence for set-based concurrent
engineering. IEEE Transactions on Engineering Management, 43(2): 165-178.
Lin, T. Y. and Stotesbury, S. D. (1988). Structural concepts and systems for architects
and engineers. Van Nostrand Reinhold Co., New York, second edition, xiii, 507 pp.,
appendix 81 pp.
Lynch, J. R and Loh, K. J. (2006). A summary review of wireless sensors and sensor
networks for structural health monitoring. The Shock and Vibration Digest, 38(2):91128.
MacGregor, J. G. and Wight, J. K. (2005). Reinforced concrete : mechanics and design.
Prentice Hall, Upper Saddle River, NJ, fourth edition, xx, 1132 pp.
Mackie, K. R. (2004). Fragility-based seismic decision making for highway overpass
bridges. PhD thesis, University of California, Berkeley.
Mackie, K. R., Wong, J.-M., and Stojadinovic, B. (2007). Comparison of post-earthquake

216
highway bridge repair costs. In Proc. Structures Congress 2007, Long Beach, CA.
ASCE/SEI.
Mackie, K. R., Wong, J.-M., and Stojadinovic, B. (2008a). Integrated probabilistic performance-based evaluation of benchmark reinforced concrete bridges. Technical Report
2007/09, University of California, Berkeley, Pacific Earthquake Engineering Research
Center, Berkeley CA.
Mackie, K. R., Wong, J.-M., and Stojadinovic, B. (2008b). Post-earthquake bridge repair
cost evaluation methodology. In Proc. Sixth National Seismic Conference on Bridges
and Highways (6NSC), Charleston, SC.
Mackie, K. R., Wong, J.-M., and Stojadinovic, B. (2008c). Probabilistic methodologies for
prediction of post-earthquake bridge repair costs and repair times. In Proc. 14th World
Conference on Earthquake Engineering (14WCEE), Beijing, China.
Miranda, E. and Ruiz-Garcia, J. (2001). Evaluation of approximate methods to estimate
maximum inelastic displacement demands. Earthquake Engineering and Structural Dynamics, 31(3):539-560.
Moehle, J., Stojadinovic, B., Der Kiureghian, A., and Yang, T. Y. (2005). An application of
PEER performance-based earthquake engineering methodology. Research Digest 20051, Pacific Earthquake Engineering Research Center.

217

Naaman, A. E. (2004). Prestressed concrete analysis and design : fundamentals. Techno


Press 3000, Ann Arbor, MI, second edition. 1072 pp.
Naeim, E, Alimoradi, A., Hagie, S., and Comartin, C. (2007). PACT : Performance Assessment Calculation Tool user guide. Technical report, Applied Technology Council
ATC-58 Project, Federal Emergency Management Agency.
Nye, W. T. and Tits, A. L. (1986). An application-oriented, optimization-based methodology for interactive design of engineering systems. International Journal of Control,
43(6):1693-1721.
Parrish, K., Wong, J.-M., Tommelein, I. D., and Stojadinovic, B. (2007). Exploration of setbased design for reinforced concrete structures. In Pasquire, C. L. and Tzortzopoulos, P.,
editors, Proc. 15th Annual Conference of the International Group for Lean Construction
(IGLC 15), East Lansing, MI.
Parrish, K., Wong, J.-M., Tommelein, I. D., and Stojadinovic, B. (2008a). Set-based design: Case study on innovative hospital design. In Tzortzopoulos, P. and Kagioglou, M.,
editors, Proc. 16th Annual Conference of the International Group for Lean Construction
(IGLC 16), Manchester, United Kingdom.
Parrish, K., Wong, J.-M., Tommelein, I. D., and Stojadinovic, B. (2008b). Value propositions for set-based design of reinforced concrete structures. In Tzortzopoulos, P. and
Kagioglou, M., editors, Proc. 16th Annual Conference of the International Group for
Lean Construction (IGLC 16), Manchester, United Kingdom.

218
Ramberg, B. and Gjesdal, K. (2005). Hermeneutics. In Zalta, E. N., editor, The Stanford
Encyclopedia of Philosophy. Stanford University, Winter 2005 edition, h t t p : / / p l a t o .
stanford.edu/archives/win2005/entries/hermeneutics/.
Saaty, T. L. (1990). How to make a decision : the Analytic Hierarchy Process. European
Journal of Operational Research, 48:9-26.
SAC Joint Venture Partnership (2000). Recommended seismic design criteria for new steel
moment-frame buildings. Technical Report FEMA-350, Federal Emergency Management Agency.
Saxon, R. (2005). Be valuable : a guide to creating value in the built environment. Constructing Excellence, London. 50 pp.
Schon, D. A. (1983). The reflective practitioner. Basic Books, New York, x, 374 pp.
Schrader, S., Riggs, W. M., and Smith, R. P. (1993). Choice over uncertainty and ambiguity
in technical problem solving. Journal of Engineering and Technology Management,
10:73-99.
Segui, W. T. (2007). Steel design. Thomson Nelson, Toronto, fourth edition.
Seismic Structural Design Associates (2005). The SlottedWebsteel moment connection.
http://www.slottedweb.com/ssda.pdf.

Shafer, C. (1985). Excellence in Teaching with the Seven Laws : A Contemporary Abridgment of Gregory's Seven Laws of Teaching. Baker Book House, Grand Rapids, MI.

219
Sobek, II, D. K., Ward, A., and Liker, J. K. (1999). Toyota's principles of set-based concurrent engineering. Sloan Management Review, 40(2):67-83.
Sonin, A. A. (2004). A generalization of the ri-theorem and dimensional analysis. Proeedings of the National Academy of Sciences of the United States of America (PNAS),
101(23):8525-8526.
Steward, D. V. (1981). The Design Structure System: A method for managing the design of
complex systems. IEEE Transactions on Engineering Management, EM-28(3):71-74.
Stojadinovic, B., Goethals, J., Wong, J.-M., Ohrui, S., Takahashi, M., Fukushima, I., and
Yamamoto, Y. (2004). Framework for integration and visualization of structural state
data. CUREE Publication CKV-03, CUREE-Kajima Joint Research Program Phase V.
Suhr, J. (1999). The Choosing by Advantages decisionmaking system. Quorum, Westport,
CT. x, 293 pp.
Sun, J., Manzanarez, R., and Nader, M. (2004). Suspension cable design of the new San
Francisco-Oakland Bay Bridge. Journal of Bridge Engineering, 9(1): 101-106.
Terry, M. S. (1974). Biblical hermeneutics : a treatise on the interpretation of the Old and
New Testaments. Zondervan, Grand Rapids, MI, second edition.
Terwiesch, C , Loch, C. H., and De Meyer, A. (2002). Exchanging preliminary information
in concurrent engineering: alternative coordination strategies. Organization Science,
13(4):402-419.

220
Thomas, R. L. (2002). Evangelical Hermeneutics. Kregel, Grand Rapids, MI.
Turk, Z. (2001). Phenomenologial foundations of conceptual product modelling in architecture, engineering and construction. Artificial Intelligence in Engineering, 15:83-92.
United States Navy (1962). Line of balance technology : a graphic method of industrial
programming. Technical report, Office of Naval Material, Government Printing Office,
Washington DC.
United States of America (2008). Part 100, "Reactor site criteria". In Code of Federal
Regulations, Title 10. United States Government Printing Office, Washington, DC.
van den Broek, J. A. (1940). Theory of limit design. Transactions of the American Society
of Civil Engineers, 105:638-730.
Ward, A., Liker, J. K., Cristiano, J. J., and Sobek, II, D. K. (1995). The second Toyota paradox: How delaying decisions can make better cars faster. Sloan Management Review,
36(3):43-61.
Werner, S., Taylor, C , Cho, S., Lavoie, J.-P, Huyck, C , Eitzel, C , Chung, H., and Eguchi,
R. (2006). REDARS 2: Methodology and software for seismic risk analysis of highway
systems. Technical Report MCEER-06-SP08, Multidisciplinary Center for Earthquake
Engineering Research (MCEER), Buffalo, NY.
Wing, J. M. (2006). Computational thinking. Communications of the ACM, 49(3):33-35.

221
Winograd, T. and Flores, F. (1986). Understanding computers and cognition : a new
foundation for design. Ablex, Norwood, NJ. xiv, 207 pp.
Wong, J.-M., Goethals, J., and Stojadinovic, B. (2005). Wireless sensor seismic response
monitoring system implemented on top of NEESgrid. Proceedings ofSPIE, 5768:74-84.
Wong, J.-M., Mackie, K. R., and Stojadinovic, B. (2006). Updating bridge fragility curves
for serviceability and extreme limit states using sensor data. In Proc. 5th National Seismic Conference (5NSC), San Mateo, CA.
Wong, J.-M., Mackie, K. R., and Stojadinovic, B. (2008). Reinforced concrete bridge seismic damage and loss scenarios. In Proc. Sixth National Seismic Conference on Bridges
and Highways, Charleston, SC.
Wong, J.-M., Parrish, K., Tommelein, I. D., and Stojadinovic, B. (2007). Communication
and process simulation of set-based design for concrete reinforcement. In Henderson,
S., Biller, B., Hsieh, M.-H., Shortle, J., Tew, J., and Barton, R., editors, Proc. of the 2007
Winter Simulation Conference (WSC07), pages 2057-2065.
Wong, J.-M. and Stojadinovic, B. (2004). Structural sensor data repository: metadata and
web-based user interface. In Proc. 13th World Conference on Earthquake Engineering,
Vancouver, B.C., Canada. Paper No. 956.
Wong, J.-M. and Stojadinovic, B. (2005). Metadata and network API aspects of a frame-

222
work for storing and retrieving civil infrastructure monitoring data. Proceedings ofSPIE,
5765:948-959.
Wong, J.-M. and Stojadinovic, B. (2006). Improving probabilistic seismic performance
evaluation using sensors and bayesian updating. In Proc. 8th National Conference on
Earthquake Engineering (8NCEE), San Francisco, CA.
Yu, L. and Kumar, A. V. (2001). An object-oriented modular framework for implementing
the finite element method. Computers and Structures, 79(9):919-928.
Yuan, M. (2008). Dynamics GIS: recognizing the dynamic nature of reality. ArcNews,
30(1): 1,4-5.

223

Appendix A
Pattern summary
A.l

Communication patterns

Narrative (4.2.1). Describe preliminary conceptual information using word descriptions instead of using sketches and drawings. Using words alone focuses on communicating functional, material, and abstract information without prematurely communicating spatial information.
Sketch (4.2.2). Describe preliminary conceptual information using sketches in order
to show spatial information and relationships between objects without implying an
artificially higher degree of precision.
Annotation (4.2.3). Display information in a visual format consisting of a main
graphic element with text that explains the important components of the graphic.
Spreadsheet (4.2.4). Present information in a tabular format in order to provide easy

224
access to data that is linked or sorted in a particular way using text and numbers.
Mechanics Model (4.2.5). Use an abstract model of a whole structure, or structure
components, based on mechanics theory to show the structure's response to loads
and deformations.
Rebar placing drawings (4.2.6). Provide drawings with information necessary for a
rebar placer to tie and place reinforcing bars within concrete formwork.
Database (4.2.7). Store structure information in a database so that it can be reused by
a computer program for further analysis.
Schedule (4.2.8). Display process information with an emphasis on showing actions
linked to time so that concurrency and dependency can be easily seen.

.2 Analysis patterns
Abstraction (5.1.1). Create a variable that is itself non-physical in order to make general analysis feasible. Convert the real spatial and material information into another
form to facilitate thinking.
Aggregation (5.1.2). Group similar items by their material or spatial information,
and treat conceptually as a single item. Simplifies a conceptual or mechanics model
by reducing the number of pieces.
Agglomeration (5.1.3). Group different types of items in a structure by their function
in order to facilitate analysis.

225
Updating (5.1.4). Use when an early initial guess needs to be updated after receiving
additional design information.
Engineering Performance (5.1.5). Measure structural design performance using the
results of a mechanics model related to the stress and strain, force and displacement
of a structural model.
Social Performance (5.1.6). Measure structural design performance using information about social conditions.

A.3 Decision patterns


Pick Any (5.2.1). Choose any option from a set of options, then confirm whether it
works.
Minimum Weight (5.2.2). Choose the design option that provides the least material
weight.
Minimum Cost (5.2.3). Choose a design option that will produce the lowest overall
cost.
Optimization (5.2.4). Choose a design option based on some optimal value of multiple performance measures involving tradeoffs between them.
Prescription (5.2.5). Follow an established rule for making a design decision, or
confirming that a candidate design choice is valid.

226
Set Exploration (5.2.6). Investigate the consequences of multiple different options
before continuing with a single option.

S-ar putea să vă placă și