You are on page 1of 10

Engineering Structures 22 (2000) 106–115

www.elsevier.com/locate/engstruct

LRFD: implementing structural reliability in professional practice


*
Bruce R. Ellingwood
Department of Civil Engineering, The Johns Hopkins University, Baltimore, MD 21218, USA

Received 1 August 1997; accepted 9 January 1998

Abstract

Structural reliability methods provide the framework for addressing building safety and serviceability issues in a rational manner.
Load and resistance factor design (LRFD) represents the first attempt in the United States to implement a rational probabilistic
approach to managing uncertainties in the building process in a structural code. The development of LRFD during the period 1969–
85, under the leadership of Professor Theodore V. Galambos, is a paradigm of collaboration between structural engineers, reliability
analysts, and the building industry. Not surprisingly, LRFD has opened the door to new research opportunities and challenges. 
1999 Elsevier Science Ltd. All rights reserved.

Keywords: Buildings (codes); Design (buildings); Reliability; Steel; Structural engineering

1. Introduction ment. The evolution of the flying buttress in gothic


cathedrals in the 12th and 13th centuries illustrates the
Structural codes provide a set of minimum technical adaptive nature of structural engineering in construction.
requirements for safe and acceptable design, a legal basis Recognizable structural engineering calculations date
for the practice of structural engineering, and a vehicle from the time of Coulomb and Navier in the late 18th
for the transfer of technology from research to practice. century. In the late 20th century, with significant
The framers of our design codes and standards, on which advances in structural engineering and with the wide-
the structural engineer relies, must address the question: spread availability of the computer as an analysis tool,
“How safe is safe enough?” on behalf of society as a the behavior of complex systems can be determined
whole. Code development is an enormous undertaking, quite accurately for design purposes. Unfortunately,
and for the most part has been done well; structural fail- despite these advances, structural loads and material
ures are rare. When failures occur, however, the conse- strengths remain unpredictable, and this uncertainty is
quences for the building occupant or user often are at the root of the structural safety problem in building
tragic, public scrutiny is intense, and economic and legal construction. Uncertainty gives rise to “risk”, rep-
consequences to the architect, engineer and owner are resenting the likelihood of an unfavorable event and its
severe. consequences in human and economic terms. Risk is
With such high stakes, it perhaps is surprising that present in all human endeavors. The primary purpose of
until recently there was no systematic way of determin- structural codes and standards is to manage and control
ing whether a structure was “safe enough”. Structural risk to socially acceptable levels.
engineering as a profession has evolved gradually over
the centuries, from a trade into a profession. During
much of that time, builders took mainly a trial-and-error 2. Structural safety
approach to safety. This learning process occurred
slowly and no doubt was accompanied by a great deal Uncertainties arise from both the demands placed on
of human misery, building loss, and personal embarrass- a system and from the capacity of the system to respond
to and withstand those demands. In many engineering
fields, particularly those involving mass-produced manu-
* Tel.: 001-410-516-8443; Fax: 001-410-516-7473; E-mail: factured products, the demands and capacities are rela-
bre@jhu.edu tively predictable; the technology is easily controlled;

0141-0296/00/$ - see front matter  1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 1 - 0 2 9 6 ( 9 8 ) 0 0 0 9 9 - 6
B.R. Ellingwood / Engineering Structures 22 (2000) 106–115 107

and ample supporting data on product performance are ened expectations on the part of the public and its scru-
available from component testing. Product design, devel- tiny of the process by which we set our standards for
opment and improvement can involve a heavy dose of public health and safety, and the litigious nature of mod-
trial and error, and acceptable performance can be ern society all have made it clear that this judgmental
ensured by the rapid feedback that occurs in the design approach to assuring building safety is becoming
process. The consequences of failure are mainly incon- increasingly unacceptable [3]. It is getting harder and
venience or economic loss. In civil/structural engineer- more costly to take corrective action following the
ing, and in other engineering fields that are professional consequences of an error in engineering judgement.
service-oriented rather than product-oriented, the situ- Many factors that determine the performance of engi-
ation is vastly different. Most structures are not mass- neered systems, while unpredictable, exhibit statistical
produced, so performance data are not easily developed regularity. This suggests that probability and statistics
under repeatable circumstances. Demands on the system can provide tools to treat these sources of uncertainty
from occupants and from natural phenomena hazards rationally in structural design. The field of structural
such as wind and earthquake are highly variable. Events reliability has developed over the past 40 years to pro-
of concern (design-basis events) are rare. Moreover, fail- vide a theoretical framework for addressing safety and
ures are highly visible and their consequences are severe. for analyzing uncertainty.
Structural engineers traditionally have approached the
problems posed by uncertainty and risk by applying a
judgmental “factor of safety” to structural calculations. 3. Role of probability in setting design-basis events
One might, for example, estimate a wind load conserva-
tively and then design the structure for a wind load that Probabilistic modeling and statistical analysis have
is twice as great; or (as in the now-archaic allowable been used extensively to determine extreme environmen-
stress design method) limit the elastic stress to some tal events due to wind, snow, earthquakes or floods for
fraction of the “failure” stress. Such factors of safety on design purposes for several decades. The alternatives of
yield strength of steel tension members have decreased choosing events that are based on a maximum historical
from about 2.5 in the 1880s to about 1.7 100 years later, occurrence or on a postulated maximum possible occur-
a decrease that presumably is associated with our rence have clear disadvantages. Maximum historical
increasing confidence in the design and construction pro- events are likely to be exceeded in the future, parti-
cess [1]. Until recently, little thought was given to what cularly if the historical record is relatively short.
the likelihood of failure of such designs might be; they Maximum possible events are based on a concurrence of
were simply judged to be “acceptably low”. It is easier conservative deterministic assumptions, and often lead to
to design a safe structure than to compute its likelihood excessive and uneconomical design requirements. The
of failure. use of probability and statistics allows some of the arbi-
As might be inferred from the previous discussion, trariness in event specification to be reduced, if not elim-
judgmental approaches to managing building risk work inated entirely.
reasonably well as long as building technology is stable, The use of statistics to analyze extreme windspeeds
there is ample opportunity to learn from experience, and dates back to at least the 1930s [4]. An early ASCE
societal expectations do not change. When structural report on wind forces on structures [5] suggested that
design is based on rare events or when new technology the use of 50-year and 100-year mean recurrence interval
is involved, however, judgement and intuition may not (MRI) windspeeds (windspeeds having probabilities of
be sufficient. Such an approach may be uneconomical in 0.02 and 0.01, respectively, of being exceeded in a given
some instances; designing a structure with stronger year) gives the designer a tool for “conservative” and
beams and columns usually (but not always) makes it “more conservative” design, based on hazard to life or
safer, but also increases its cost. Perhaps more seriously, property from a structural failure. Probability in earth-
it may fail to account for different (perhaps higher) per- quake engineering found a place in the specification of
formance categories required for certain facilities. For the seismic hazard [6], expressed in terms of an effective
example, the U.S. Nuclear Regulatory Commission peak ground acceleration, spectral acceleration at speci-
strives to limit the probability of radionuclide release fied periods, or similar parameters. In the mid-1970s, a
from a nuclear power plant accident to less than one in probabilistic seismic hazard study was undertaken [7],
a million per year [2]. The fact that we now have several ultimately leading to a seismic hazard map with contours
thousand years of accumulated plant operating experi- of ground acceleration with 10% probability of being
ence furnishes only anecdotal evidence of whether cur- exceeded in 50 years (equivalently, a 475-year MRI). As
rent requirements for design, construction, operation and with wind, no technical justification was provided for
maintenance are sufficient to meet this goal. The after- these numbers; the 50-year interval was selected for con-
math of recent natural and man-made disasters, the rapid venience, while the 10% probability was judged to be
evolution of design and construction methods, height- “reasonable and conservative”.
108 B.R. Ellingwood / Engineering Structures 22 (2000) 106–115

A review of other contemporary literature indicates in which FR(x) ⫽ cumulative distribution function of R
little justification for the quantitative probability levels and fQ(x) ⫽ density function of Q. This equation is
selected to specify design-basis events in modern struc- impractical for design purposes, as anyone who ever has
tural codes. A 50- to 100-year MRI event is not incon- worked with it can appreciate. For one thing, it requires
sistent with the largest (or near-largest) event that an a knowledge of the probability laws for R and Q, which
engineer might expect to experience during his or her may vary for different structural actions and limit states.
professional career. On the other hand, probabilistic Also, there is the problem of how to handle the numeri-
models of load events and statistical methods for treating cal integration in the typical iterative design context.
data do provide a rational framework for processing Thus, prior to the late 1960s, most of the problems
observed data consistently and for extrapolating beyond addressed in the reliability literature were primarily
the range of observation. Thus, as a result of studies such theoretical and relatively simple in nature [9].
as those above, probabilistic methods have come to be Within the period 1968–72, however, the scope of
accepted in structural codes during the past two decades structural reliability expanded rapidly beyond a rela-
in setting design levels for natural phenomena hazards. tively narrow research community to a broader group of
The ASCE 7 Standard [8] has required that ordinary engineers concerned with improving the standard devel-
building structures be designed for windspeeds or opment process. Within this period, we have the First
ground snow loads with MRIs of 50 years since 1972; International Conference on Structural Safety and
buildings have been required to withstand earthquake Reliability [10] (held, appropriately enough, at the
ground shaking corresponding to a MRI of 475 years Smithsonian Institution in Washington, DC); a collection
since 1982 (reportedly, this may increase to 2500 years of papers on practical aspects of structural safety pub-
in the next edition for certain structures); dams are lished in the ACI Journal in the Sept–Dec 1969 issues,
designed to withstand 100-year flood levels, and so on. followed by a technical session organized by ACI on
These criteria address the uncertainty in the design event probabilistic design of concrete buildings in 1971; and
with a frequency distribution that describes it explicitly an ASCE specialty conference on reliability of metal
with a probabilistic model. By setting the design level buildings [11]. Also in that period, efforts were initiated
at some small probability of being exceeded, some meas- to transform Eq. (1) into something more practical for
ure of risk consistency is introduced that is not possible design [12–14]. It was a remarkably quick transition
with a purely judgemental safety factor or a “maximum from academic to applied research, and did not occur
event” specification. by happenstance.
The above approach, of course, deals only with the The first moves toward what we now call limit states
demand portion of the structural design problem. Inte- design (plastic design for steel or strength design for
grating demand and capacity into a reliability-based concrete) had preceded the expansion to practical appli-
structural design code or specification requires a higher cations in the reliability area by only about a decade.
level of sophistication, and advances in this arena have Limit states design, in contrast to allowable stress design
been more recent. The remainder of this paper presents (ASD), requires careful thought about how best to cope
a personal view of how structural reliability has evolved with possible modes of structural behavior (elastic ver-
in meeting the needs of the standard development com- sus nonlinear response), analysis techniques (first-order
munity, and will identify some of the challenges facing versus second-order), and performance limits
us, now that first-generation reliability-based structural (serviceability, onset of structural nonlinearity, incipient
design codes are in place. system instability). ASD does not deal with these issues
in an entirely rational manner when it deals with them
at all. More often, the complexities raised by these issues
4. Probability-based limit states design simply are shrouded in the overall factor of safety that
is used to reduce the critical failure stress (yielding,
The use of structural reliability theory as a tool in buckling, fracture) down into the elastic range, where it
structural code development in the United States can be can be accommodated in a first-order elastic structural
traced back to the late 1960s. In classical reliability analysis. Leaders in the structural engineering profession
theory, the structural action(s) due to the applied load(s), who wrestled with these issues during this period real-
Q, and the resistance, R, are modeled by random vari- ized early on that limit states design would require a
ables. In the simplest representation, failure occurs if R more rational way of dealing with uncertainties [1].
is less than Q. The probability of this event is, Although many people contributed to the debate, the
contributions of George Winter of Cornell University,
who was one of the eminent individuals who served on



both the AISC Specification Advisory Committee and
P[Fail] ⫽ Pf ⫽ FR(x)fQ(x)dx (1) the ACI Committee 318 during those years, are
0 especially noteworthy. Professor Winter observed that a
B.R. Ellingwood / Engineering Structures 22 (2000) 106–115 109

more rational treatment of the uncertainties in loads and AISC Specification Committee, first at a week-long
strengths was essential and, in addition, argued strongly meeting in August 1981, and subsequently at yearly
for the development of common load requirements for meetings over a period of several years. The first LRFD
different construction materials. Comments from this specification was published in 1986 and the second in
segment of the standards community found fruit in the 1994 [17]. By the end of the decade all steel design
reliability community, and is the main reason why the improvements and specification revisions will be tailored
papers in these conferences and publications during primarily for LRFD.
1968–72 had, for the first time, a strong practical current The plea for common load requirements for different
running though them. construction materials was answered only after the first
One notable example of this move toward practical draft of the LRFD specification was completed. The
implementation can be found in the notion of the 1970s saw a number of advances in first-order reliability
reliability index, ␤, as an alternative (to Pf) measure of analysis, stochastic load modeling, and supporting data
reliability. The reliability index was introduced in an acquisition. These advances made it possible to develop
attempt to avoid conceptual and practical difficulties a set of common load requirements, including load fac-
(problems of computation, lack of data, modeling errors) tors and load combinations, for inclusion in ANSI Stan-
with the use of Eq. (1). In its initial implementation, ␤ dard A58 (now ASCE Standard 7) in a relatively short
was evaluated simply as a function of the means and period in 1979. The procedure by which this was done—
standard deviations (or coefficient of variations) of collection and analysis of statistical data on loads and
resistance and structural actions. (Later, techniques for strengths, reliability assessment of structural members
incorporating information on probability distributions for which it was believed that current design was accept-
were developed, leading ultimately to what we now call able, selection of reliability targets, and finally setting
first-order reliability methods.) Following the simple load criteria to meet those targets—is documented in two
form of Eq. (1), if the margin of safety, M, is defined articles appearing in May 1982 [18,19]. These load cri-
as M ⫽ R ⫺ Q, with mean E[M] and standard deviation teria were adopted by ballot for ANSI Standard A58.1-
SD[M], then ␤ ⫽ E[M]/SD[M]. For common structural 1982; they have been reaffirmed and thus have remained
engineering practice, this number usually is on the order essentially unchanged (with the exception of the earth-
of 2 to 6; many engineers at the time were more comfort- quake load factor) since then [8].
able with such numbers than with probabilities of 10−5 LRFD successfully transformed the classical structural
or less (in the case where R and Q are normal, ␤ ⫽ reliability equation into a form that was practical for
⌽−1(1 ⫺ Pf). Moreover, ␤ could be computed with only design, even by engineers not familiar with the reliability
a knowledge of “most likely value” (reflected, approxi- concepts. From the performance requirement that the
mately, in E[M]) and “uncertainty” (reflected in SD[M]), member reliability should exceed the target reliability
quantities with some familiarity, rather than the entire index, ␤, Eq. (1) was replaced by the familiar LRFD for-
distribution, which was an unfamiliar concept. Since ␤ mat,
incorporated uncertainty explicitly, it seemed to offer an
improvement over the traditional factor of safety, where ␾Rn ⫽ Σ␥iQni (2)
there was abundant evidence that uncertainty was not
treated rationally. in which Rn and Qni are the nominal strengths and loads,
Load and Resistance Factor Design (LRFD) rep- respectively, traditionally provided in familiar codes and
resents the first attempt in the United States to implement standards, and ␾ and ␥i are resistance and load factors
rational probabilistic thinking in the context of a modern that reflect (a) uncertainty in strength and load, and (b)
limit states structural code. The American Iron and Steel consequences of failure, measured by the associated
Institute (AISI) and the American Institute of Steel Con- value of ␤. The left hand side of Eq. (2) is the purview
struction (AISC) initiated a research project in 1969 to of AISC (or other material specification); the right hand
develop a practical steel design specification using side is defined by ASCE Standard 7. The specific load
reliability principles to treat uncertainties in loads and factors—1.2 on dead load, 1.6 on live load, 0.9 on yield-
in structural strength. The project was under the direc- ing in tension, and so forth—are familiar to most readers.
tion of Professor Theodore V. Galambos, who was A similar approach is followed by our colleagues in Can-
assisted by Dr M.K. Ravindra, and was guided by an ada, Mexico, Europe and Japan. It may seem like a small
advisory committee of experts in steel structures, struc- step, but it was not; the move toward LRFD required a
tural design, and reliability theory. The major portion of thorough re-examination of the philosophical under-
the technical work was performed during the period pinnings of ASD (or lack thereof) and led to changes in
1969–76. The technical basis for LRFD is explained in the procedures by which Rn is calculated for different
a collection of eight papers published in September 1978 limit states as well.
[15,16]. A period of trial design and refinement fol- The development of the first LRFD specification for
lowed. The draft LRFD specification was debated by the steel structures from 1969 to 1985 provides the paradigm
110 B.R. Ellingwood / Engineering Structures 22 (2000) 106–115

for a collaboration of those in reliability research, a parti- from ASD to LRFD. In my view, the competitive press-
cular construction technology, and professional practice ures of the international marketplace will ensure that
to work together to improve the process by which build- advances in computerized structural analysis and design
ing structures are engineered. In retrospect, the process will continue unabated. Those who fail to adapt to this
leading to the LRFD Specification may seem inordi- reality will find themselves out of business early in the
nately long. However, it was a learning process that 21st century.
involved a broad group of professionals with diverse Structural system effects are reflected in the current
backgrounds and expertise. A more recent effort to reliability-based codes, but only indirectly. Examples of
develop ASCE Standard 16-95 on LRFD for engineered this include the K-factor used in column design and the
wood construction was essentially completed during the R and Cd factors used to account for nonlinear behavior
period 1988–94, or 6 years. The NCHRP project to in the context of an elastic analysis in earthquake-resist-
develop an AASHTO LRFD Code for highway bridge ant design [20]. Improved structural modeling (both lin-
design was accomplished in a similar time frame. ear and nonlinear) made possible by computational
These probability-based code development efforts advances will enable us to address the issue of designing
were successful, in my view, because they were man- the structural system as a system rather than as a collec-
aged by engineering researchers who were familiar, first tion of members. The need for such an approach is evi-
of all, with the structural engineering issues involved, dent in several areas, including stability analysis of tall
and who were willing to adopt the reliability tools neces- buildings, but perhaps most immediately in earthquake
sary for analyzing uncertainty. It is problematic whether engineering, where most recent advances (including
these code development efforts would have been suc- Chapter 9 of ASCE Standard 7-95 and AISC’s Seismic
cessful had their leadership come from the reliability Provisions for Structural Steel Buildings) have been
research community. On the other hand, the reliability aimed at ensuring that the system can tolerate or with-
analysts played an absolutely essential role in providing stand strong ground motions without life-threatening
the means for dealing with the multitude of uncertainties damage. The performance of special moment-resisting
attendant to a structural code. Such opportunities for steel frames during the Northridge Earthquake is evi-
synergism are replete in modern structural engineering. dence of not only the benefits of steel construction (no
building collapses occurred) but the need for more accur-
ate models of connection behavior.
5. The future of reliability in structural Redundancy and load redistribution subsequent to
engineering initial yielding were not taken into account explicitly in
the structural reliability assessment underlying LRFD.
The next decade presents exciting opportunities for This can lead to a false impression of the reliability of
structural engineers, standard writers, and reliability ana- even an indeterminate beam, let alone a more complex
lysts to continue to work together to rationalize the structural system. To illustrate this point, consider the
building design process. The following is a collection, two-span continuous beam in Fig. 1, designed for the
in no particular order and certainly not all-inclusive, of nominal loads shown. Assume that a floor slab provides
some of the challenges facing us. Readers no doubt will full lateral support, and that the beam is compact; thus,
add to the list. LTB or local instability need not be considered. We will
consider two design situations: (1) ASD (elastic) design;
5.1. Advances in computation and systems analysis (2) LRFD Section F1.1 (first plastic hinge). Using LRFD
Section F1.2d (plastic design) leads to the same required
Advances in computation will play a key role in meet- section as (2) because the positive moment within the
ing many of the challenges below. Such advances have left span controls design. Table 1 summarizes the steel
been occurred exponentially during the last 28 years beam sections obtained by these design procedures. We
since the work on reliability-based design standards such note that LRFD permits approximately an 11% reduction
as LRFD was initiated. Numerous structural analysis in beam weight from ASD.
programs are now available for performing first-order Next, two possible beam limit states are considered
analysis, second-order nonlinear analysis, and so forth. for purposes of reliability assessment: first yield and
They are efficient in terms of resources required, and plastic collapse (other intermediate limit states may be
often are accompanied by elegant user-oriented inter- of interest in some applications). The fact that dead and
faces and design aids. They offer the prospect of introd- live loads, wD and wL, both are random variables means
ucing reliability-based design directly in the design that the location of the maximum positive moment in
office. Of course, they do require a human capital invest- the left span is random. This source of uncertainty has
ment to be fully useful as productivity tools. One objec- a relatively small impact on the object of this particular
tion to LRFD, often voiced in professional and trade example or the results, and will be neglected for sim-
association venues, centers on the cost of transitioning plicity. These two limit states then are (the divisor 12
B.R. Ellingwood / Engineering Structures 22 (2000) 106–115 111

Fig. 1. Continuous beam design by ASD and LRFD.

Table 1 lapse. On the other hand, when design is completely


Design of two-span continuous beam computer-automated and every beam and column is
Criterion Section Reliability index, ␤ designed at or close to the limit state or when the struc-
ture has little capability for redistributing force following
First yield Collapse first yield or failure, there may be little additional margin
for error. The reliability of such structural systems can
ASD W21 ⫻ 62 2.43 3.03 be substantially less. How much less has not yet been
LRFD F1.1 W24 ⫻ 55 2.02 2.76
thoroughly investigated, and requires additional con-
sideration in the next generation of LRFD specifications.
Closely coupled to the issue of system behavior is that
enables wD and wL to be expressed in units of kips/ft of performance-based design. It has been somewhat sur-
while Fy, Sx and Zx are in customary units): prising and amusing to me to see performance-based
design recently introduced in several quarters as a new
Yield: BFySx/12 ⫺ (62.5 wD ⫹ 85.9375wL) ⫽ 0 (3a) concept. The performance concept was introduced in
1970 in the Operation Breakthrough Project, sponsored
Plastic: BFyZx/12 ⫺ 77.2068(wD ⫹ wL) ⫽ 0 (3b) by HUD at the National Bureau of Standards, and a set
of (primitive) reliability-based criteria for strength and
in which Fy ⫽ yield strength; Sx and Zx ⫽ elastic and serviceability limit states were published in 1977 [21].
plastic section moduli, respectively. The probabilistic At that time, unfortunately, computational resources
models for these terms are available elsewhere [18,19]. were insufficient to cope with some of the problems of
Using customary first-order reliability analyis, the system design-by-analysis, and the alternative of validat-
reliability indices, ␤, (based on the maximum structural ing designs by structural testing was viewed as too
action in a 50-year service period) are computed for first expensive. Twenty years later, resources now are avail-
yield and plastic collapse and are presented in the final able to take full advantage of the performance concept.
two columns of Table 1. There is a strong move in this direction in the SAC Steel
Several points are worthy of note. For one, the differ- Project being funded by FEMA, and the Department of
ence between the “yield” and “collapse” columns, in Energy Standard [22] on design and evaluation for natu-
terms of limit state probability, is about a factor of 7. ral phenomena hazards has already adopted it.
This increase in reliability against collapse above first Encapsulated in the performance concept are the
yield reflects redundancy in the beam; the difference notions of identifying different performance limits
would have been only minor had the beam (structure) explicitly (e.g., impaired service function; loss of occu-
been statically determinate. For another, the use of stan- pancy; life-threatening damage), relating structural
dard hot-rolled shapes leads to sections greater than or response (perhaps nonlinear) to a level commensurate
equal to the required section; in the examples in Table with the performance limit identified (nonstructural dam-
1 this oversizing ranged from 3 to 8%. As a result, the age, first yielding or hinge, incipient instability), and
␤’s in the final column of Table 1 fall above the mini- associating with each a specific numerical risk goal,
mum of approximately 2.5 (with respect to a 50-year expressed in terms of probability. One might depict this
service period), on which the LRFD Specification [17] sort of behavior by a family of “fragility” curves,
is based. Moreover, in most structural framing systems, describing the (conditional) probability of failure of the
columns are extended over several stories or the same system, for a particular demand. This is illustrated in
beam section is used for several bays for ease in fabri- Fig. 2, where a family of such curves are presented for
cation, even if the full section capacity is not required. a frame; the vertical axis plots conditional limit state
This additional token design conservatism increases the probability, P[LS|W ⫽ w], while demand (measured here
actual reliability further over that indicated by code cali- in terms of lateral force, w) is measured on the horizontal
brations (on which the minimums provided by the code axis. The slope of each fragility is a measure of the
are based), and make it possible to tolerate some local uncertainty in the structural behavior for that limit con-
damage in the structure without fear of structural col- dition. The separation or clustering of these fragilities at
112 B.R. Ellingwood / Engineering Structures 22 (2000) 106–115

Fig. 2. Fragility models for performance limits.

a given fractile, say 0.05, conveys an idea of the robust- side our technical realm where we can influence it and
ness of the system; curves that are tightly clustered are into the political arena.
indicative of a system that has little capacity beyond the
first sign of damage. As a specific example for earth- 5.2. Target reliability measures
quake-resistant design, one might use fragilities to differ-
entiate between the design requirements for different cat- Probability-based analysis and design concepts are
egories of building occupancy based on seismic risk founded on the notion of a “target” reliability or risk as
[23]. Since the variability in the seismic hazard might a quantitative measure of acceptable performance. Sel-
be strongly site-dependent, such an approach may also ecting reliability targets for structural engineering pur-
require designing for different levels of the earthquake poses is difficult because of the unique nature of the
hazard (expressed, e.g., in terms of spectral facility in question, limited data, and small limit state
accelerations) to meet all stated building performance probabilities of relevance. The reliability index, ␤ as a
goals. Moreover, the load criteria may very well be site- measure of reliability in LRFD was adopted, in part, to
dependent. Some evidence of the need for this differen- avoid the problem of specifying a numerical value of
tiation can also be found in the wind area, where recent risk. Calculated risks or failure probabilities that are not
research [24] suggests that wind risk is substantially consistent with what has been observed in practice create
higher in hurricane-prone areas than in regions affected a credibility problem for those using reliability tools in
mainly by strong extratropical winds. structural analysis and design. Those of us who are
System reliability analysis techniques must be working toward advancing reliability-based design must
developed to provide the numerical risks to support such develop a fuller appreciation of this fact and be willing
design and decision-making. Such techniques can range to communicate the risk concepts to our colleagues and
from the complex to the simple, but many are now others in terms that are easily understood.
within reach. A performance-based approach to struc- One might attempt to establish reliability targets by
tural design seems complex when compared to the for- considering risks that arise from other exposures and
mat and content of current structural standards. Nonethe- human activities. For example, the risk of death from
less, it is an idea worth considering for the next disease provides a psychological yardstick, of sorts,
generation of reliability-based standards and for LRFD. against which the acceptability of other risks might be
In the light of building performance during recent natural measured. It is well to have some sense of what these
disasters, it seems likely that more informative measures risks, expressed here in terms of annual frequencies, are
of building risk may be required by the insurance indus- in the United States [1,25,26]. Overall mortality risk
try as a condition for underwriting policies for new or (uncorrected for exposure) is about 8.6 ⫻ 10−3/year; due
existing building construction. It would be well for the to cancer and heart disease, 5.7 ⫻ 10−3/year. Individual
structural engineering profession to get a handle on this risk of death due to motor vehicle accidents is about 2
before subsequent events move the decision process out- ⫻ 10−4/year; due to home accidents, 1 ⫻ 10−4; due to
B.R. Ellingwood / Engineering Structures 22 (2000) 106–115 113

fire, 1.5 ⫻ 10−5. Homicide risk is about 1.5 ⫻ 10−4/year, 5.3. Designing for balanced risks
taken over the whole population; in our inner cities, it
is higher by an order of magnitude. Commercial aviation It seems self-evident that engineering design should
risk to a well-traveled professor of structural reliability strive to achieve some balance among competing risks
is about 5 ⫻ 10−5/year. Risks that are incurred volun- by focusing efforts on reducing the worst of them. Costly
tarily can be about 100 times those incurred involuntarily efforts can be misdirected toward reducing one source
and still be acceptable [27,28]. of risk to vanishingly small levels, leaving other sources
In the civil engineering field, the observation that, on unaddressed. We have one example of this in the build-
average, 1% of our dams suffer a failure of some kind ing code community, manifested by the perennial argu-
within 20 years of construction implies a dam failure rate ment concerning research funding in the wind versus
of about 5 ⫻ 10−4/year. Bridge collapse probabilities are earthquake engineering areas. In the nuclear power area,
on the order of 10−4/year. Most estimates of building significant resources were directed for a number of years
structural collapse probability are in the range 10−6/year toward large LOCA mitigation until PRAs demonstrated
to 10−5/year. that this was only a minor source of offsite risk in com-
What, then, are we to make of the “risk” associated parison to other sources, such as loss of offsite power.
with LRFD, and its target reliability indices of about Reliability methods can show where current code devel-
2.5–3.0 [29], obtained by calibrating to traditional design opment activities are not in tune with the risks. The first
practice for members in which the limit state is yield in generation of reliability-based codes does not take full
tension or first plastic hinge in flexure? The correspond- advantage of this capability.
ing limit state probability (for general localized yielding) As a simple example, suppose that structural failures
would be on the order of 0.0025 in 50 years or, approxi- can result from any one of M mutually exclusive and
mately, 5 ⫻ 10−5/year, assuming that the variability in collectively exhaustive hazards, Hi, each of which occurs
the strength is small in comparison with that in the struc- (annually) with probability P[Hi]. Suppose, further, that
the consequences of failure due to each Hi are compara-
tural action due to the loads. The structural collapse
ble (this assumption is not necessary in decision analysis,
probability would be less, depending on the extent to
but allows the basic concept to be presented in a math-
which the structural system permits load redistribution
ematically simple form). The risk of failure (measured
following first yield. In a truss of 100 members, in which
in this simple example by the probability, Pf) of a build-
each member is sized exactly at the LRFD limit, we
ing would then be,
would expect to see one or two of the members at incipi-


ent yield sometime during a service life of 50 years. M
These probabilities are not inconsistent with some of the Pf ⫽ P[Fail|Hi]P[Hi] (4)
societal risks enumerated above. Perhaps they can be i⫽1
explained and justified more easily than we have been
led to believe. in which P[Fail|Hi] ⫽ (conditional) probability of fail-
Determining acceptable risk hinges on socio-political ure, given the occurrence of Hi. If each of the terms
as well as technical issues and concerns. While risks are P[Fail|Hi] P[Hi] is of comparable magnitude, then the
real, the attitudes of the average person toward them are risk is “balanced.” Conversely, if one of these product
not grounded in reality. People are more willing to terms is orders of magnitude higher than the others, then
accept risks if they understand them. Public reaction to socio-technical efforts to address the other sources of
nuclear power is a good example of what happens when hazard will have only a marginal impact on overall risk
people are not well-informed about a particular tech- and might be criticized as a waste of public resources.
nology and its associated risk, and a few highly vocal A general strategy for mitigating risk from diverse
special interests move into the void and steer the national sources should examine each of the terms in Eq. (4), in
debate. Finding solutions to the problem of setting turn, and would wring out as much benefit as possible
acceptable risks in structural engineering will always be from each. It might be possible to reduce risk by con-
difficult given the rarity of design-basis events. Engin- trolling or eliminating the hazard, which would reduce
eers must communicate risk information and voice their P[Hi]. This might not be socially acceptable in some
views to their peers in profession and in the social arena. instances (e.g., natural gas cannot be proscribed to con-
We cannot back away from the acceptable risk issue by trol the possibility of explosions in apartments because
talking about notional failure probabilities, reliability gas is necessary for heating and cooking). In others, it
indices, and such. More precise ways of expressing might not be possible (e.g., we cannot control the occur-
reliability targets will have to be used to take full advan- rence of hurricanes or tornadoes). The structural stan-
tage in LRFD of concurrent advances in structural analy- dards community at present focuses its attention entirely
sis and reliability analysis. on the terms P[Fail|Hi], striving to reduce the effect of
the hazard by making the conditional probability
114 B.R. Ellingwood / Engineering Structures 22 (2000) 106–115

acceptably low. However, we only pay attention to a few iod of previous service provides evidence of minimum
of the terms in this summation. It is little wonder that integrity. The proposed rehabilitation or retrofit may
our calculated probabilities do not agree well with obser- arise from a change of building occupancy or usage, and
vations. the reliability goals for the structure may have changed.
There is a need to address the performance of struc- Nondestructive evaluation plays a key role in periodic
tures against hazards other than those for which the condition assessment; there is uncertainty in all common
building was explicitly designed. These so-called abnor- NDE technologies, both in locating flaws and in measur-
mal load events would include structurally significant ing their potential impact on structural behavior. More-
fires, explosions of natural gas, detonation of high over, there is uncertainty associated with various mainte-
explosives, or vehicular collisions, events with inci- nance and repair actions. These uncertainties must be
dences typically on the order of 10−6 to 10−4 dwelling taken into account in reliability-based condition assess-
unit/year [1]. Tragically, it is becoming more necessary ments of existing buildings. The current load and resist-
to take such a view, in order to mitigate the occurrences ance criteria in LRFD, which were intended for design-
of events such as those at the World Trade Center in ing new buildings and were developed from time-
New York and the Murrah Federal Building in Oklah- invariant probabilistic models of strength in which the
oma City. Probability-based LRFD criteria for fire-resist- above sources of uncertainty were not represented, may
ant structural design are within reach [30], and special not apply for evaluating an existing building [31]. While
load requirements can be found in the Commentary C2.5 quantitative risk or margins assessments have been pro-
to ASCE Standard 7-95. posed for some critical facilities [22], such assessments
may be difficult if implemented for routine civil con-
5.4. Safety of buildings under construction struction. Developing LRFD tools for evaluating existing
steel structures presents a challenge.
The LRFD Specification and ASCE Standard 7-95 do
not address buildings under construction. Reliability 5.6. Human error
analysis for buildings during construction is a difficult
problem, made more so by the nearly infinite number of After the structural reliability analyses are completed,
ways that building constuction can proceed. Contractors we are left with the fact that most spectacular and cata-
maintain their competitive edge by ingenious construc- strophic structural failures have occurred as the result
tion sequencing and likely would view any standard with of human error, not stochastic variation in the material
suspicion, as possibly inhibiting their flexibility. An strengths or loads that are reflected in the load and resist-
ASCE Standard Committee is developing a construction ance criteria in LRFD. Errors are distributed throughout
load standard. Incorporation of reliability concepts has the building process: in design, construction, and service
proved difficult. life. Human error caused the collapse of the Skyline
Tower in Fairfax, VA in 1973; the Hartford, CT Civic
5.5. Condition assessment of existing structures Arena Roof in 1978; the Willow Island Cooling Tower
in WVA in 1978; the elevated walkway at the Hyatt
Like any other man-made product, buildings wear out Regency Hotel in Kansas City, MO in 1981; the L’Am-
over a period of many years from the accumulated biance Plaza apartments in Bridgeport, CT in 1987. Con-
effects of use and the environment. Even a casual glance trary to intuition, design errors cause a significant frac-
at the decaying infrastructure in our largest cities high- tion of such failures. Steel structures are not immune to
lights the importance of structural aging as a national such failures. Previous research has indicated that the
problem. Most structures are not designed for a finite contribution of the human error term(s) in the summation
service life. Criteria are needed to evaluate existing in Eq. (4) may be on the order of 5 to 10 times the
structures—to perform condition assessments of their magnitude of the term(s) associated with stochastic
fitness for future service. It may be surprising to some variability in loads and strengths. Research is needed to
to learn that the LRFD Specification has no provisions identify error sources and to develop strategies for their
for load testing existing buildings; those engineers who control in design, construction and operation. Ignoring
are asked to specify a load test often must consult Sec- the possibility of such failures in risk assessment is
tion 20 of ACI Standard 318 for reinforced concrete likely to lead to an overly optimistic view of facility
buildings, where some rudimentary provisions are found. safety.
The time-variant aspect of the reliability analysis is
particularly significant for condition assessment of exist-
ing structures, since their strength and stiffness change 6. A few parting thoughts
over time as a result of aging effects that are uncertain
and difficult to model analytically. On the other hand, Structural reliability methods furnish code developers
the fact that the structure has survived an extended per- with quantitative bases for engineering decision-making
B.R. Ellingwood / Engineering Structures 22 (2000) 106–115 115

and a clear audit trail for the treatment of uncertainties maximum acceleration in rock in the contiguous United States.
in the building process. They do not remove the burden Open File Report 76-416, US Geological Survey, Golden, CO,
1976.
of making the decisions, however. This frustrates some [8] Minimum design loads for buildings and other structures (ASCE
engineers who, when first exposed to these methods, had 7-95). American Society of Civil Engineers, New York, 1996.
hoped to more or less automate the decision process. [9] Freudenthal A, Garrelts J, Shinozuka M. The analysis of struc-
Structural reliability provides a forum for a dialog tural safety. J Struct Div ASCE 1966;92(1):267–325.
between the research and practice ends of structural [10] Freudenthal AM, editor. Proceedings, International Conference
on Structural Safety and Reliability. Oxford: Pergamon Press,
engineering on how best to manage uncertainty and risk 1972.
in structural design. This dialog must establish the link [11] Safety and reliability of metal structures. Committee on Metals,
between performance goals embodied in LRFD and American Society of Civil Engineers, New York, 1972.
society’s willingness to accept risk. Thus, we may expect [12] Cornell CA. A probability-based structural code. ACI Journal
to see structural reliability have an increasing role in the 1969;66(12):974–85.
[13] Rosenblueth E, Esteva L. Reliability basis for some Mexican
code development process, enabling structural engineers codes. In: Probabilistic design of reinforced concrete buildings.
to reduce or otherwise manage risk, to address newly ACI SP-31, American Concrete Institute, Detroit, MI, 1972.
defined hazards, and to identify and respond to new chal- [14] Ang AH-S, Cornell CA. Reliability bases of structural safety and
lenges in building design. design. J Struct Div ASCE 1974;100(9):1755–69.
[15] Ravindra MK, Galambos TV. Load and resistance factor design
for steel. J Struct Div ASCE 1978;104(9):1337–54.
[16] Galambos TV, Ravindra MK. Properties for steel for use in
Acknowledgements LRFD. J Struct Div ASCE 1978;104(9):1459–68.
[17] Load and resistance factor design specification for structural steel
buildings. American Institute of Steel Contruction, Chicago, IL,
This paper was presented in a shortened form at a 1993.
Symposium on Innovations in Structural Design: [18] Galambos TV. Probability-based load criteria: assessment of cur-
Strength, Stability, Reliability, which was held in Min- rent design practice. J Struct Div ASCE 1982;108(5):959–77.
neapolis, MN, June 1997, to honor Theodore V. Gal- [19] Ellingwood B. Probability-based load criteria: load factors and
load combinations. J Struct Div ASCE 1982;108(5):978–97.
ambos. Its preparation was supported, in part, by [20] Uang C-M. Establishing R (or Rw) and Cd factors in for building
Lockheed-Martin Energy Corporation under Award seismic provisions. J Struct Engrg ASCE 1991;117(1):19–28.
19X-SP638V and by the US Army Corps of Engineers [21] Performance criteria resource document for innovative construc-
under Award DACW39-95-0018. This support is grate- tion. NBSIR 77-1316, National Bureau of Standards, Wash-
fully acknowledged. However, all views herein are those ington, DC, 1977.
[22] Natural phenomena hazards design and evaluation criteria for
of the author, and do not necessarily reflect the views department of energy facilities (DOE-STD-1020-94). US Depart-
of the above organizations. ment of Energy, Washington, DC, 1994.
[23] Cornell CA. Reliability-based earthquake-resistant design: the
future. Presented at Emilio Rosenblueth Symposium, 11th World
Conf on Earthquake Engrg, Acapulco, Mexico, June 1996.
References [24] Simiu E, Heckert NA, Whalen T. Estimates of hurricane wind
speeds by the “peaks over threshold” method. NIST Technical
[1] Ellingwood B. Probability-based codified design: past Note 1416, National Institute of Standards and Technology, Gai-
accomplishments and future challenges. Struct Safety thersburg, MD, 1996.
1994;13(3):159–76. [25] Statistical abstract of the United States. US Department of Com-
[2] Okrent D. The safety goals of the U.S. Nuclear Regulatory Com- merce, Washington, DC, 1995.
mission. Science 1987;236:296–300. [26] Wilson R, Crouch EAC. Risk assessment and comparisons: an
[3] Committee on Risk Appraisal in the Development of Facilities introduction. Science 1987;236:267–70.
Design Criteria. Uses of risk analysis to achieve balanced safety [27] Starr C. Social benefits versus technological risk. Science
in building design and operations. Building Research Board, 1969;168:1232–8.
Washington, DC: National Academy Press, 1991. [28] Pate-Cornell E. Quantitative safety goals for risk management of
[4] Davenport AG. A rationale for the determination of design wind industrial facilities. Struct Safety 1994;13(3):145–57.
velocities. J Struct Div ASCE 1960;86(5):39–68. [29] Ellingwood B, Galambos TV. Probability-based criteria for struc-
[5] Wind forces on structures. Transactions, ASCE tural design. Struct Safety 1982;1(1):15–26.
1961;126(PartII):1124–98. [30] Ellingwood B, Corotis RB. Load combinations for buildings
[6] Cornell CA. Engineering seismic risk analysis. Bul Seismological exposed to fires. Engineering Journal, AISC 1991;28(1):37–44.
Soc Am 1968;58(5):1583–606. [31] Ellingwood B. Reliability-based condition assessment and LRFD
[7] Algermissen ST, Perkins DM. A probabilistic estimate of for existing structures. Struct Safety 1996;18(2):67–80.

You might also like