Vous êtes sur la page 1sur 23

The Shock and Vibration Digest

http://svd.sagepub.com Impacts of Epistemic (Bias) Uncertainty on Structural Identification of Constructed (Civil) Systems
F. L. Moon and A. E. Aktan The Shock and Vibration Digest 2006; 38; 399 DOI: 10.1177/0583102406068068 The online version of this article can be found at: http://svd.sagepub.com/cgi/content/abstract/38/5/399

Published by:
http://www.sagepublications.com

Additional services and information for The Shock and Vibration Digest can be found at: Email Alerts: http://svd.sagepub.com/cgi/alerts Subscriptions: http://svd.sagepub.com/subscriptions Reprints: http://www.sagepub.com/journalsReprints.nav Permissions: http://www.sagepub.com/journalsPermissions.nav Citations (this article cites 100 articles hosted on the SAGE Journals Online and HighWire Press platforms): http://svd.sagepub.com/cgi/content/refs/38/5/399

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

Articles

Impacts of Epistemic (Bias) Uncertainty on Structural Identification of Constructed (Civil) Systems


F.L. Moon and A.E. Aktan
ABSTRACTThe primary goal of this paper is to discuss the state-of-the-art related to the structural identification (St-ID) of constructed systems, and to point out the most pressing needs for research and applications. As a subset of the system identification concept, St-ID aims to develop representative mathematical models of manufactured (mechanical) and constructed (civil) structural systems through the correlation of experimentally measured inputs and outputs. While the concept of St-ID has matured over the past three decades and it is widely and reliably applied to manufactured systems, St-ID for constructed systems remains in its infancy and has enjoyed only sparse implementation in practice, and generally only for signature structures. The authors believe that this delayed progress is principally the result of a lack of appreciation for the inherent distinctions between constructed systems and their manufactured counterparts. Through this paper the authors hope to clearly convey the unique challenges associated with the St-ID of constructed systems and motivate researchers from engineering mechanics to join in the investigation of these issues and to ultimately aid in advancing the art of St-ID of constructed systems. KEYWORDS: structural identification, constructed systems, epistemic uncertainty, aleatory uncertainty

adopted. The crux of this view is that the establishment of reality requires that mathematical models be hypothesized and validated based on observations of the physical world. According to Eves (1990), this concept dates back to Plato, who claimed that the reality which scientific thought is seeking must be expressible in mathematical terms, mathematics being the most precise and definite kind of thinking of which we are capable. These ideas embody the scientific method and also provide the foundation for the concept of Sys-ID. In its most general form, Sys-ID aims to establish the relationship between the Physical, Mental and PlatonicMathematical worlds. To provide a mathematical definition of Sys-ID, consider a dynamic mechanical system (Zhang et al., 2004). The SysID process begins by hypothesizing a mathematical representation of the system, such as a state-space model: x(t) = F(x(), < t; u(), t) y(t) = G(x(), t; v(), t) (1) (2)

1. Introduction 1.1. Origin and Definition of System Identification The system identification (Sys-ID) concept, defined as the estimation of a system based on the correlation of inputs and outputs, originated in electrical engineering in relation to circuit and control theory. Over the past few decades, the SysID concept has flourished as numerous engineering disciplines have recognized its value. Today, Sys-ID serves as a fundamental prerequisite for addressing systems problems in mathematics, physics, economics, social sciences, and throughout engineering (Kossiakoff and Sweet, 2002). To illustrate the origins of Sys-ID, the view put forth by Penrose (2005) that reality is composed of three worlds Physical, Mental and Platonic-Mathematical (Figure 1) is
F.L. Moon, Department of Civil, Architectural and Environmental Engineering, Drexel University, Philadelphia, PA. A.E. Aktan, Drexel Intelligent Infrastructure and Transportation Safety Institute, Drexel University, Philadelphia, PA The Shock and Vibration Digest, Vol. 38, No. 5, September 2006 399420 2006 SAGE Publications DOI: 10.1177/0583102406068068

where x(t) represents the various states of the system, u(t) denotes the external inputs to the system, y(t) denotes the outputs of the system, and v(t) denotes observation uncertainty (all at time instant t), F( ) is the functional relation of state transition, and G( ) is the functional relation of outputs

Figure 1. Three worlds (adapted from Penrose 2005).

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

400

The Shock and Vibration Digest / September 2006

and states. In general, the states which define the system are not directly observable. As a result, Sys-ID aims to infer the states x(t) from the output y(t), and in some cases the input u(t) and measurement uncertainty v(t), by finding an appropriate estimator H( ), such that: x(t) = H(y(), t; u(), t; v(), t) 1.2. Structural Identification Structural identification (St-ID) is a subset of Sys-ID which has been adapted to mechanical (manufactured) and civil (constructed) structural systems, and may be defined as the parametric correlation of structural response characteristics predicted by a mathematical model with analogous quantities derived from experimental measurements (Doebling et al., 2000). This paradigm was first introduced to engineering mechanics researchers by Hart and Yao (1977) and to civil-structural engineering researchers by Liu and Yao (1978). These seminal papers have inspired many researchers to investigate various aspects of St-ID, and nearly 30 years later St-ID remains an active research area in both engineering mechanics and civil-structural engineering. Figure 2 provides an illustration of the St-ID process, which is a specialization of the general Sys-ID concept. Perhaps the most relevant distinction between St-ID applications is related to the type of structural system (manufactured or constructed) that is being identified. In the case of manufactured systems, the concept of St-ID has matured over the past three decades and is widely and reliably applied to various systems. In contrast, while hundreds of investigations into the St-ID of constructed systems have been performed (see the Appendix), the use of St-ID for constructed systems remains in its infancy and has enjoyed only sparse implementation in practice, and (generally) only for signature structures. While several researchers have argued this is (3)

primarily due to a lack of practical sensing and networking technology, recent advances in these areas have not been accompanied by widespread implementation of St-ID. Rather, it is becoming increasingly clear that the lack of implementation (and also the skepticism towards St-ID held by many owners/stewards of constructed systems) stems from an inability to interpret measurements reliably enough to influence management decisions. This is compounded by the problem that in many cases irrelevant and unreliable data, especially erroneous identification of deterioration or damage, become a liability for managers. The authors believe that this difficulty principally results from a lack of appreciation for the inherent distinctions between constructed systems and their manufactured counterparts. Overcoming this challenge requires that the recent advances in St-ID, control and monitoring of manufactured systems be adapted to explicitly recognize and address the unique attributes of constructed systems. An example of this type of development is the recent progress associated with operational (or output-only) modal analysis, which recognizes the cost and difficulty of performing forced vibration tests on large constructed systems (see, for example, Brownjohn et al., 1992; Fujino et al., 1999; Wenzel and Pichler, 2005). In addition, several researchers have developed St-ID approaches that explicitly address aleatory uncertainty (uncertainty due to natural randomness), which can be significant for constructed systems (Bucher et al., 2003; Yuen and Katafygiotis, 2002; Beck and Katafygiotis, 1998; Beck, 1990). However, while these advances are relevant, the distinctions they address are far from exhaustive, and thus a wider, more sustained effort is necessary. In particular, the challenges associated with epistemic uncertainty (due to our inability to correctly understand, model or predict the behavior of constructed systems) have to be recognized and addressed. As a first step, civil-structural engineers must identify the principal distinctions that lead to the unique challenges associated with the St-ID of constructed systems and communicate them to the wider community. 1.3. Scope and Objectives To that end, the goal of this paper is to discuss the stateof-the-art related to the St-ID of constructed systems, and point out the most pressing needs for research and applications. The authors hope to clearly convey the unique challenges associated with the St-ID of constructed systems and motivate researchers from engineering mechanics to join in the investigation of these issues and ultimately aid in advancing the state-of-the-art. The paper begins with a general discussion of the primary distinctions between manufactured and constructed systems. This is followed by a description of suggested definitions and classifications related to St-ID that may help initiate discussions in the community with the aim of establishing a universal nomenclature in this area. To illustrate the significant potential of St-ID to provide knowledge related to the performance of constructed systems, an overview of a few selected studies is presented. Next various challenges that continue to hinder the reliable implementation of current St-ID approaches are discussed and illustrated through the use of past studies. The paper concludes with a discussion of the authors vision of the direction of future research on this topic.

Figure 2. Structural Identification Stages.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS

401

2. Distinctions Between Manufactured and Constructed Systems The authors believe that the most fundamental distinction between the St-ID of manufactured and constructed systems is related to the level of uncertainty, both aleatory and epistemic, that must be managed (Ang and De Leon, 2005). While the general definition of Sys-ID (Equations 13) does recognize measurement uncertainty, this uncertainty has traditionally been taken as random noise. In the case of constructed systems, however, measured data and a priori models are also frequently subjected to significant epistemic uncertainty, which may be deterministic, and is often many times larger than the uncertainty due to natural randomness. Table 1 provides some unique attributes of constructed systems that have been observed during St-ID studies over the past decades (see the Appendix). While attributes contributing to aleatory uncertainty (including heterogeneity, sensor/ cable noise, etc.) have been widely recognized, the considerable epistemic uncertainty that affects response measurements of constructed systems has not. Consider that most of the attributes listed in Table 1 violate common fundamental assumptions of many St-ID techniques, i.e., that structures
Table 1. Some of the unique attributes of constructed systems. Heterogeneity

are linear, stationary, and observable. Depending on the severity with which these implicit assumptions are violated, epistemic uncertainty is introduced into the results. We note that various researchers have investigated non-physics-based (NPB) approaches to St-ID to mitigate these challenges. However, these approaches are not without their shortcomings, as the nature of their output does not easily facilitate the use of heuristic knowledge and thus considerable uncertainty may be associated with the interpretation and decision-making processes. In addition to the increased complexities listed in Table 1, constructed systems (unlike manufactured systems) cannot be isolated from sources of uncertainty during the St-ID process. For example, a Boeing 747 can be removed from service and tested with controlled boundary conditions in a controlled laboratory environment. In contrast, because of their size and fusion with their operating environment (e.g., soil-structure-interaction, temperature, etc.), constructed systems must be tested with their in-service, nebulous boundary conditions and in their non-stationary environments. To compound this type of direct uncertainty, the St-ID of a constructed system is also plagued by many sources of indirect uncertainty. For example, human constraints imposed by owners/stewards,

Materials, member proportions, detailing, etc. can vary considerably from member to member, and within a member. Deterioration and damage compounds these variations and makes discretization difficult and sometimes unmanageable without heuristics. Constructed systems have unobservable soil-foundation interfaces that are often non-stationary in their contact properties. Soil and even rock properties change with pressure, moisture, temperature and time. Most constructed systems, and especially bridge systems, are designed with movement systems and/or force releases. These systems are usually unobservable and behave differently under different levels of force and temperature. Constructed systems have many types of local, regional and global/external redundancies. These redundancies are highly affected by temperature changes and temperature gradients, which results in intrinsic forces and changes in element properties. Constructed systems maintain complex and non-stationary intrinsic forces due to dead weight, construction loads/staging, temperature effects, deterioration, damage, overloads, etc. These intrinsic forces are nearly impossible to measure in an absolute sense and even their changes often overwhelm the forces due to transient live loads. Element, connection and global behavior of real constructed systems exhibit many different types of nonlinearity that change at different limit-states. Cracking, material yielding, local instability, connection slip, interface friction, etc. are all associated with both hardening and softening behaviors that are also frequently visco-elastic and/or visco-plastic. Constructed systems are non-stationary due to the non-stationary nature of inputs (temperature, radiation, traffic, wind, etc). Temperature and humidity effects are highly complex: changes and variations in the rate of change in ambient, regional and local temperatures and humidity of the structure and the soil may lead to intrinsic forces and also induce changes in boundary and continuity conditions. Nearly all constructed systems are custom-designed for specific applications and their mechanical characteristics are strongly affected by events during and immediately following their construction. While types of constructed systems may be grouped based on their primary structural system, size, materials, etc., applying results from a single structure to a larger population of structures is challenging due to their inherent uniqueness. Constructed systems such as major highway bridges or combinations of several bridges and tunnels within regional transportation networks may be several miles or more in length, cost several billions of dollars (~15 Billion for the Big Dig in Boston), and be expected to remain in service for well over 100 years. The size and lifecycle impedes our ability to view such systems in a holistic manner over a sufficient proportion of their lifecycles and further compounds the natural variability and uncertainty in their mechanical characteristics.

Boundaries Continuity

Redundancy

Intrinsic forces

Types of nonlinearity

Non-stationary nature

Uniqueness

Geometric, temporal scale, cost, lifecycle

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

402

The Shock and Vibration Digest / September 2006

Table 2. Classification of Analytical Modeling Forms Physics-Based Models Continua Models Theory of Elasticity Field and Wave Eqns Mechanics of Materials Models - Bernoulli - Vlasov, etc. Discrete Models Spring-Mass-Damper K,M,C Coefficients Smeared-Macro or Element Level FE Models Micro-level, Geometric-Replica FE Models Modal Models: - Modal Parameters - Ritz Vectors Non-Physics-Based Models Semantic Models Ontologies Semiotic Models Meta Models - Input-Output - Rule-based Probabilistic Models Bootstrap Standard Prob. Distributions (Monte Carlo) Numerical Models Agents ARMA, ARMAX, Rational Polynomial ANN Signal/Pattern Analysis, Wavelet, EMD, others

research funding mechanisms, legal issues, and user demands severely limit both the number of studies, and the rigor with which and limit states at which they can be performed. As a result, St-ID studies to date have been performed in something of an ad hoc manner aimed at gaining detailed information about specific structures. Due to the uniqueness of constructed systems and the targeted aim of these studies, only a very limited amount of generalized knowledge about the behavior of such systems has been developed. 3. Classifications of Analytical Models and Experimental Methods As apparent from Figure 2, St-ID requires the integration of analytical, experimental and information technologies. Although information technology (IT) is already embedded in both the analytical and experimental steps, the execution of Steps 4 and 5 may require specialized IT applications related to data visualization, archival and warehousing of the multi-scale, multi-modal and multi-bandwidth data. Experimental technologies include excitation systems, those related to the measuring and monitoring of local and global response quantities, and equipment for nondestructive evaluation (NDE) techniques. Analytical technologies include forward and reverse applications of computer-aided drafting (CAD), analytical modeling, and various forms of linear and nonlinear analysis including numerical and computational aspects as well as pre- and post-processing tool requirements. While all aspects of the St-ID process are critical to the overall success of the effort, it is important to recognize that analytical modeling is perhaps the most significant step. By the time an engineer has conceptualized and constructed an analytical model of a constructed system, the utility of the StID effort is largely defined. The following sections provide a brief discussion of common analytical modeling approaches (both physics-based and non-physics-based; see Table 2) and experimental technologies that have been developed or employed for the St-ID of constructed systems. 3.1. Physics-based Approaches Since the 1970s numerous researchers have investigated the use of physics-based (PB) St-ID approaches to actual

constructed systems (see, e.g., Hornbuckle et al., 1973; Douglas and Reid, 1982; Natke and Yao, 1986, 1989; Brownjohn et al., 1987, 2003; Biswas et al., 1989; Agbabian et al., 1990; Stubbs et al., 1992; Kou and DeWolf, 1997; Doebling et al., 1998; Farrar and Doebling, 1998; Fujino and Abe, 2002; Teughels and De Roeck, 2004; Maeck and De Roeck, 2003; and Aktan et al., 1997a, 1998a). The primary benefit of PB approaches is that the identified model facilitates the use of heuristics and can be used to explicitly simulate behavior under various critical loading conditions. While several researchers have investigated the use of nonlinear models (Jayakumar and Beck, 1988; Chassiakos et al., 1995; Kapania and Park, 1997; Smyth et al., 1999; Naghavi and Aktan, 2003), currently the most commonly employed PB St-ID approach relies on modal analysis algorithms and linear(ized) finite element (FE) models. Depending on how severely the structure being identified violates the implicit assumptions of this approach (i.e., linearity, stationarity, and observability), significant errors can occur. Worse still, unless independent experimental approaches are employed (e.g., comparison of static and modal flexibility) it may be very difficult to assess whether an identification has been sufficiently complete and reliable. Over the past decade, there has been a growing awareness of the importance of incorporating uncertainty within the St-ID process for constructed systems. In response, several researchers have developed methods that explicitly incorporate the uncertainties associated with the identified modal parameters within the model updating process (Bucher et al., 2003; Yuen and Katafygiotis, 2002; Beck and Katafygiotis, 1998; Beck, 1990). In addition, several tools have been developed for system identification with special emphasis on uncertainty analysis and quantification. For example, tools such as NESSUS (developed by the Southwest Research Corporation [www.swri.org]) and DAKOTA (developed by Sandia National Laboratories [endo.sandia.gov/DAKOTA]) can be used in conjunction with commercially available FE packages to quantify uncertainty or perform sensitivity analyses, among other capabilities. However, while these developments are quite relevant, they are geared towards addressing aleatory uncertainty and do not provide insight into the effects of epistemic uncertainty (e.g., the appropriateness of a priori models). To the authors knowledge, there has not been any

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS

403

systematic attempt to investigate the impact of epistemic uncertainties on PB St-ID approaches. 3.2. Non-physics-based Approaches The most common non-physics based (NPB) St-ID techniques employed in civil and mechanical engineering include artificial neural networks (ANNs) (Chang et al., 2001; Masri et al., 1996; Nakamura et al., 1998; Zapico et al., 2003), wavelet decomposition (Al-Khalidy et al., 1997; Gurley and Kareem, 1999; Hou et al., 2000), auto-regressive moving average vector (ARMAV) models (Shinozuka and Ghanem, 1995; Andersen and Kirkegaard, 1998; Bodeux and Golinval 2001), and empirical mode decomposition (EMD) in conjunction with the Hilbert-Huang Transform (Huang et al., 1998; Vincent et al., 1999; Yang et al., 2001). In addition, Table 2 includes some more novel forms of NPB models such as agents, input-output, ontology and semiotic models that are beginning to find use for modeling complex systems such as infrastructures involving human, natural and engineered elements. We would also note that there are many innovative approaches to NPB modeling in medicine, economics, social sciences, communications, etc. The main advantage of the NPB techniques is that they are data-driven, i.e., the construction of NPB models is dependent entirely on the data provided. This data driven nature makes them attractive for modeling complex phenomena, automation, real-time St-ID, continuous monitoring, and minimizing errors caused by user interaction. While these benefits cannot be ignored, it is equally important to recognize that they may not be effective in the absence of PB models and heuristics. While NPB approaches may be able to track and identify changes or detect outliers in a structures response, relating this change to its root mechanism is not possible with NPB models alone. For example, consider the non-stationary behavior of constructed systems caused by temperature fluctuations. While several NPB models have successfully captured this phenomenon (Sohn et al., 1999; Zhao and DeWolf, 2000; Peeters et al., 2001) they were not able to identify the exact mechanisms, i.e., whether ambient temperature, temperature gradients within the structure, temperature shocks or the effect of temperature on boundary conditions and movement systems, etc., were the root causes of the changes in mechanical characteristics. Thus NPB models cannot, in isolation, distinguish natural changes in the structural state from damage. 3.3. Experimental Technologies Most of the common experimental technologies that have been explored for St-ID are classified in Table 3 (Aktan et al., 2002). Measuring and monitoring geometry (local or global), nondestructive evaluation (NDE) for characterizing localized conditions and/or material properties, short-term and long-term structural testing (monitoring) are first differentiated by whether measurements are focused on capturing local or global properties, and by the duration of measurements. Specific experimental techniques that are listed under each broader class of experiment require highly specialized skills, knowledge and experience. The common elements intersecting all of the experimental techniques are sensing, data acquisition, communication and computing.

We note that various classes of experiments (such as modal analysis) and many of the specific technologies within each class have developed along distinct paths under the auspices of professional organizations such as ASTM, ASNDT, SEM, SPIE, IMAC, various TRB Committees, AWS, etc. However, classification of the entire spectrum of experimental technologies and how these may be integrated for synergistic applications for the St-ID of a constructed system have not yet been addressed by any organization. Further, the challenges in integrating experiment, analysis and information technology within the framework of St-ID are not universally acknowledged or addressed. We note that there is no unique procedure for developing and calibrating an experimental technology. Various NDE tools have been developed by empirical approaches based on statistical regression, such as the impact hammer for measuring in-situ concrete properties. However, as a technology becomes more complex, the importance of first carrying out a rigorous theoretical formulation and a sensitivity analysis cannot be ignored. In the case of complex NDE probes whose fundamentals are best understood by physics experts, the success of a technology becomes contingent upon the success of interdisciplinary collaborations between scientists who are knowledgeable about the theoretical basis of the technology and the application engineers who establish the application parameters. Even once this challenge is overcome, there still remain further challenges related to prototyping, commercialization and proper marketing and implementations of a technology. Similar issues exist for global structural testing approaches such as modal analysis and instrumented monitoring. Although these technologies have been extensively used for mechanical systems such as automobiles and airplanes, their application to constructed systems requires technology transfer research and calibration through large numbers of applications. In addition, many civil engineers are unaware of the requirements and challenges related to obtaining reliable data and are led to believe that making measurements is only a matter of purchasing and installing a sensor. As a result, hands-on training that addresses the formulation of experimental goals, design of experiments, sensor selection/implementation, and, most importantly, interpretation of measured data, is also needed. 4. Motivations for the St-Id of Constructed Systems While the recent advent of the structural health monitoring (SHM) paradigm, coupled with innovations and advances in related technologies, has greatly increased interest in St-ID, it is interesting to note that applying this concept to constructed systems is far from new. In fact, some of the most prolific structural engineers since the industrial revolution have recognized and exploited some form of St-ID to aid in their understanding of structural performance. For example, Robert Maillart argued in 1907 that the testing of completed structures should be widespread and is required to conceptualize three-dimensional behavior and to check design assumptions through comparing deflections (Billington, 1983). Many technical committees within the ASCE, ACI and AISC, as well as many workshops and conferences held by geotechnical and structural engineers, have reiterated the importance of

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

404

The Shock and Vibration Digest / September 2006

Table 3. Classification of Experimental Approaches Local Short-Term Structural Testing NDE Load Testing (Quasi-Static Testing) Material Testing Thermal Magnetic Ultrasound Acoustic Electrical ElectroChemical Optical Crawlspeed Truck Loading Special Loading Devices Input by Traffic Static Truck Loading Input by Traffic WeighIn-Motion
*

Global Long-Term Testing with Intermittent or Continuous Monitoring

Vibration Testing (Dynamic Testing) Controlled Measure Input & Output Uncontrolled Measure Output Only

Controlled Measure Input & Output

Uncontrolled Measure Output Only Measure Input by WIM* & Output

Low-Bandwidth Measurements Construction Effects Wind/Ambient Weather Conditions (Temperature, Humidity, etc.) Force, Deformation, Strain, etc. Geometry

High-Bandwidth Measurements Vibrations Truck Loads Security Monitoring Operations/Tolling Traffic Incidents, Accidents, Fire Impacts Earthquake

Impact ForcedVibration by Exciter Input by Traffic, Wind, etc.

controlled testing of full-scale soil-foundation-constructed systems to understand their kinematics and load distribution paths at various limit states. From the first application of StID, such studies have been motivated by our incomplete understanding of constructed system behavior and aimed at reducing such epistemic uncertainty. In this regard, civil engineers are distinguished as the oldest engineering discipline, with the least understanding of the actual mechanical characteristics, behaviors and performances of the systems they design. Unless civil engineers make an effort to better know their products, the transformation of their discipline to a performance-based one will be difficult. The following sections provide some examples that demonstrate the potential of St-ID to manage epistemic uncertainty in order to transform civil engineering practice to a performance-based one. 4.1. Interactions with Operating Environment One of the most widely investigated and reported sources of uncertainty in making measurements on constructed systems involves their observed non-stationary behavior caused by environmental inputs such as temperature and humidity. For example, past monitoring studies have identified significant fluctuations in modal frequencies and mode shapes; as large as 40% over a 24 hour period (Alampalli, 1998). However, the majority of the reports indicate impacts of around 520% (Farrar et al., 1994; Cornwell et al., 1999; Lenett, 1998; Roberts and Pearson, 1998; Rohrmann et al., 2000; Fu and DeWolf, 2001; and Peeters et al., 2001). In addition to affecting modal parameters, several researchers have shown that temperature fluctuations significantly affect the intrinsic forces and structural properties of constructed systems (Aktan, 1995; Lenett, 1998; Aktan et al., 2000a). In some cases, temperature shocks, associated with fast-moving weather fronts, have been observed to permanently alter the intrinsic force distri-

bution and move constructed systems to a new state of equilibrium (Catbas and Aktan, 2002). This implies that for some structures the variation of temperatures within the constructed system and the history of temperature changes may be more critical parameters than the current ambient temperature. While the impacts of temperature on structural behavior have long been recognized, they must still be considered sources of epistemic uncertainty, as their accurate a priori modeling remains problematic. Consider that while several monitoring studies have indicated that temperature-induced stresses on bridges may exceed those caused by permitted vehicles (Lenett, 1998), we remain unable to accurately represent them in the design, evaluation, and management process. Over the last decade, however, the use of St-ID has successfully reduced the level of epistemic uncertainty associated with damage detection due to the non-stationary nature of constructed systems. This issue was discussed by Farrar et al. (1994) in the report of forced and ambient vibration tests conducted on the I-40 Bridge over the Rio Grande. The authors noted a 510% difference in modal frequencies and mode shapes over a 24-hour period, which was attributed largely to temperature effects. Following initial testing, the bridge was intentionally damaged by introducing four levels of cuts in one of the two primary plate girders. Results indicated that only when the plate girder was cut halfway through (i.e., the largest damage case) was there an appreciable change in natural frequencies. However, even in this case the change was comparable to that attributed to temperature changes. Similarly, Peeters and De Roeck (2000), Lenett (1998), and Alampalli (1998) also report that environmental influences cause changes in modal parameters that can mask the changes caused by damage. To investigate this phenomenon, some researchers have examined the dependence of modal properties on temperature fluctuations using St-ID. One of the more interesting examples of this type of investigation was reported by

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS

405

Peeters and De Roeck (2001) and focused on the data obtained from the testing of the Z-24 Bridge. In this case, an approximately bi-linear relationship between frequency and temperature was obtained, with the intersection of the two portions occurring around 0C (Figure 3). The authors traced this bilinear dependence to the influence of the wearing surface. Specifically, during warm periods the asphalt stiffness was such that it did not contribute to the response, but during colder periods (<0C) the stiffness of the wearing surface increased and thus participated in the response of the bridge. As an aside, from Figure 3 it is clear that the large variation in eigenfrequencies observed around 0C would introduce significant uncertainty into the St-ID of the structure in this region. Other researchers have traced non-stationary behavior to stiffening of expansion bearings that introduce longitudinal forces into the bridge (Fu and DeWolf, 2001; Alampalli, 1998), temperature gradients (Cornwell et al., 1999), and internal redundancies (Catbas and Aktan, 2002; Rohrmann et al., 2000). While the root mechanisms of this non-stationarity have been identified in many cases, the complex nature of this phenomenon makes modeling its behavior with physicsbased models difficult (Catbas et al., 2003). As a result, most modeling approaches employed have been non-physics-based. Using data from the Alamosa Canyon Bridge, Sohn et al. (1999) developed and validated the effectiveness of a linear four input (two time and two spatial) adaptive filter to reproduce the non-stationary behavior of the first two modal frequencies. Using this filter, the authors were able to separate non-stationary behaviors due to temperature changes from those likely to have been caused by water absorption (increased mass) following heavy rains. In addition, Peeters et al. (2001) used ARX models to distinguish the effects of temperature fluctuations from real damage events on the modal frequencies of the Z-24 Bridge. The authors conclude that a single temperature measurement was sufficient to identify an accurate model, and that ARX models that include the thermal dynamics of the bridge are superior to simple static regression models. It should be pointed out, however, that this study only considered the linear portion of the

temperature versus eigenfrequency plots shown in Figure 3 (i.e., > 0C). 4.2. System-Level Structural Behavior Due to difficulties conducting experimental studies on complete structural systems, the vast majority of structural behavior investigations conducted over the past half century have focused on components and idealized subassemblies (e.g., beams, columns, connections). While this research has developed a wealth of knowledge about component behavior, such knowledge has proved difficult to extrapolate to the system-level and thus considerable epistemic uncertainty associated with system-level behavior still exists. This arises primarily due to system-level mechanisms that cannot be established by exclusively investigating component-level behavior, since they result from continuity conditions between the various elements and global boundary conditions of the complete structure. The effects of this uncertainty are most apparent when examining the results of destructive tests of constructed systems, which have typically revealed failure mechanisms that come as a surprise to even the most experienced investigator and system capacities that vary considerably from predictions (>100%). The following sub-sections provide examples of how St-ID has been leveraged to uncover system-level boundary and continuity condition. 4.2.1 Boundary Conditions One of the principle motivations for performing field tests of constructed systems is the significant influence actual boundary conditions can have on constructed system performance. In general, this issue must be studied in the field as it represents the interaction between constructed and natural systems. To investigate the boundary conditions of the Millikan Library Building on Cal Techs campus, Foutch et al. (1975), Luco et al. (1988), and Wong et al. (1988) performed forced vibration tests designed to isolate the effects of soil-structure interaction. Results indicated modal parameters of the structure-foundation-soil system varied considerably from those associated with fixed-base conditions. For example, it was found that the rigid-body motion associated with the translation and rotation of the base accounts for more than 30% of the total roof response. Similar observations were made by Foutch et al. (1989) during forcedvibration tests of buildings in Mexico-City following the 1985 earthquake. Aktan et al. (1995b) reported the results of quasi-static destructive tests (using hydraulic actuators with rock anchors) of an aged steel truss bridge. The authors report that as the bridge was loaded past the proportional limit, the bottom chord came into contact with the back wall of the abutment. The stiffness of the abutment restrained the horizontal displacement of the supports and resulted in an increase in vertical stiffness of the bridge due to archingaction. As a result, the authors concluded that definitions of damage that rely on global stiffness reduction are too simplistic to be applied reliably to real-life structures. Eberhard and Marsh (1997a,b) imposed quasi-static loads on a threespan continuous reinforced concrete bridge and isolated reinforced concrete bridge bents. Results indicated that approximately 90% of the lateral bridge stiffness was due to arching-action of the deck enabled by the longitudinal confinement supplied by the abutments. In addition, the testing

Figure 3. The first eigenfrequency of the Z-24 Bridge versus temperature (Peeters et al. 2001).

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

406

The Shock and Vibration Digest / September 2006

of individual bridge bents revealed ductile behavior, despite 1960-era, pre-seismic design, details. This ductility was attributed to the rotational flexibility of the soil, which alleviated high rotational demands on poorly detailed portions of the bents. 4.2.2 Continuity Conditions In addition to global boundary conditions, mechanisms that may arise due to continuity conditions between various components are a principle source of uncertainty related to system-level behavior. However, since the mechanisms that affect these conditions are contained within the structural systems, they can be investigated either in the field or in the laboratory as long as complete structural systems are employed. Perhaps the most remarkable example was observed during the US-Japan Coordinated Research Program on RC Buildings. In this research program, leading earthquake engineering researchers in Japan and the US predicted the responses of a 7-story prototype building that was designed with an idealized structural wall-frame system and was laterally loaded in the laboratory in one plane (Bertero et al., 1985). The predicted shear and overturning capacities of the building, based on pseudo-3D modeling of the building, were found to be in error by more than 300% after the tests were conducted and the actual capacities of the building were measured. Through the use of St-ID, it was revealed that mechanisms of global rocking and outrigger-action had resulted in the vast differences between the predicted and measured responses. However, the identification of these mechanisms required St-ID, as none of the earthquake engineering experts, including many of the best academic and practicing structural engineers, anticipated them. Catbas and Aktan (2002) and Lenett (1998) report the St-ID of a deteriorated steel stringer-concrete slab bridge using impact and crawl-speed truck testing procedures. Results indicated that although no shear connections were present between the slab and stringers the bridge was behaving in a composite manner due to chemical bonding. Because of this continuity, the bridge was able to withstand 1.2 times the bridge rating load with live load stresses less than 2,500 psi. Yi et al. (2006) and Moon et al. (2006) performed quasi-static tests of a full-scale unreinforced masonry (URM) building to investigate seismic performance. Results indicated that the continuity between primary in-plane components as well as the continuity between in-plane and outof-plane walls provided lateral stiffness and capacities up to 100% greater than obtained from the currently accepted analysis approach. By including the flange participation and overturning effects that resulted from the identified continuity conditions, errors between experimental and simulation results were reduced to less than 20%. 5. Uncertainties that Continue to Challenge Reliable St-Id Although St-ID is principally motivated by the need to quantify and reduce uncertainty, its reliable implementation is also plagued by uncertainty. For example, aleatory uncertainty is introduced by various sources throughout the St-ID process and must be properly addressed and quantified to provide accurate estimates of the reliability of the identified model. In addition, significant epistemic uncertainty is also introduced into the St-ID process, especially through viola-

tion of the underlying assumptions of the selected St-ID procedure. To organize the discussion of how uncertainties impact St-ID, a framework that categorizes some of the more common sources of uncertainty based on the step at which they enter into the St-ID process was assembled (Figure 4). Although perhaps outside the domain of engineering study, it is also important to recognize that the impact of these uncertainties is further compounded by human errors due to inattention, thoughtlessness, inexperience, omission or commission. To provide some tangible examples of how the uncertainties listed in Figure 4 have hampered St-ID, the following sections provide selected examples from past research studies. While each of the uncertainties listed in Figure 4 can potentially impede reliable St-ID, for the sake of brevity and due to the authors experience, the following discussion is limited to the uncertainties associated with controlled experimentation (Step 3). 5.1. Structural Complexity A primary source of epistemic uncertainty is related to the relatively high level of structural complexity typical of constructed systems. Such complexities can be classified loosely as either internal or external. For example, internal complexities that introduce significant uncertainty include high degrees of redundancy, potentially large and unknown intrinsic stresses, local nonlinearities, and shakedown phenomena. Sources of complexity such as soil-foundation-substructure-superstructure interaction as well as the interaction between the structure and its operating environment in general are classified as external complexities. Admittedly, this distinction is somewhat muddled as significant coupling between external and internal complexities clearly exist (e.g., redundancy and temperature effects). In general, these issues have affected both the identification of structures and the approaches used for interpreting response data (e.g., damage detection). The following subsections illustrate some common sources of epistemic uncertainty associated with structural complexity. 5.1.1 Intrinsic Stresses Due to many parameters, including dead weight, construction loads/staging, temperature effects, deterioration, damage, overloads, etc., constructed systems maintain complex and non-stationary intrinsic forces. While these intrinsic forces, in an absolute sense, are nearly unobservable, they are perhaps the most critical parameter affecting structural capacity as their magnitudes often overwhelm the forces due to transient live loads (Catbas and Aktan, 2002). As a result, even the most rigorous St-ID cannot be expected to produce a model that can accurately predict the response of the structure at more advanced post-damage limit states, as such predictions require accurate identification of intrinsic forces. For example, Aktan et al. (1995b; 1995c) report the results of destructive tests on two aged steel truss bridges using hydraulic actuators with rock anchors. At each successive limit state a rigorous St-ID of each bridge was carried out using data from a series of nondestructive tests that excited the bridges with a rotating weight shaker, crawl-speed truck, and a linear-mass shaker. Results indicated that due to non-uniform stress distributions and unobservable intrinsic forces, analytical models, identified at each limit state, were unable to accurately predict the bridges response at

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS

407

Figure 4. Uncertainties associated with the St-ID of constructed systems.

more advanced limit states. As a result, it is clear that considerable epistemic uncertainty exists for all identified models, due to their inability to properly capture such an influential parameter. 5.1.2 Nebulous Boundary and Continuity Conditions In many cases, constructed systems have boundary or continuity conditions that are either non-stationary or nonlinear even at operating limit states. Depending on how severe these complexities are, significant uncertainty can be introduced into the results of experimental modal analysis. For example, due to intrinsic forces, several bearings of the Seymour Bridge displayed only intermittent contact at the operating limit state (Lenett, 1998). As a result of these contact problems, the initial attempts to identify global modal parameters were unsuccessful. However, once the bridge was welded to the supports to prevent lift off and contact problems (i.e., once the supports were idealized), the identification was successful. It is important to note that the detection of erroneous modal identification results was principally facili-

tated by comparisons with the independently determined static flexibility. 5.1.3 Interaction between Primary Sub-Structural Components For constructed systems composed of structural components with very large differences in stiffness, such as the rigidly designed towers and flexible superstructures of many suspension bridges, the eigenproblem becomes weakly-coupled. As a result, the vibration modes measured from one portion of the structure may be completely unrelated to that portion of the structures mass, stiffness, and damping properties. Rather, the modes or spikes in the response may simply be a reflection (or echo) of the modes in the other portion of the structure. Grimmelsman and Aktan (2005) identified this issue during the St-ID of the Brooklyn Bridge carried out as part of a seismic assessment study focused on the masonry towers. Due to the difficulty and cost associated with providing forced excitation, the study relied on ambient excitation to obtain response data.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

408

The Shock and Vibration Digest / September 2006

Results indicated that the ambient excitation was dominated by traffic which entered the towers only after being filtered by the flexible spans. That is, the towers were essentially being excited by a small number of harmonic signals associated with the dominate modes of the spans. Obviously, this violates assumptions of broadband excitation (and spatially distributed excitation) and thus introduces considerable uncertainty into the St-ID process and the identified tower modes. As a result, utilizing the identified tower modes to provide insight into seismic response, characterized by base excitation with higher frequency content, is highly questionable. 5.2. Excitation Issues Experimental studies of constructed systems have been carried out using a wide range of excitation sources including ambient excitation sources (McLamore et al., 1971; AbdelGhaffar and Housner, 1978), test vehicles (Oehler, 1957; Billing, 1984), eccentric mass vibrators (Hudson, 1970; Kuribayashi and Iwasaki, 1973; Shepard and Charleson, 1971; Ibanez et al., 1981), linear mass shakers (Galambos and Mayes, 1978; Salane et al., 1981), step-relaxation techniques (Douglas, 1976; Buckle et al., 1986; Gundy et al., 1981), instrumented impact hammers/drop weight devices (Lee et al., 1987; Miller et al., 1992; Raghavendrachar and Aktan, 1992), rocket engines (Iwasaki et al., 1972; Srinivasan et al., 1981), and weighted pendulums (Buckland et al., 1979). While several of these excitation systems, such as impact/drop weight devices and eccentric mass vibrators, are essentially scaled up versions of excitation systems previously applied to mechanical systems, others were developed specifically for constructed systems either for convenience or to better approximate service loading (e.g., test vehicle, ambient excitation). The following sections discuss some sources of uncertainty associated with the more popular excitation methods. 5.2.1 Amplitude of Input A principal source of uncertainty is related to the amplitude of the selected excitation. For very large structures it becomes very difficult to supply sufficient forced excitation to overcome the ambient vibration. This is particularly the case for very flexible structures such as cabled-stayed or suspension bridges (Wenzel and Pichler, 2005). In addition, it is important to realize that although modal frequencies and mode shapes tend to be independent of excitation level, (up to the proportional limit of the structure) this is not the case for damping. Farrar et al. (2000) report the results of vibration tests conducted on the Alamosa Canyon Bridge in which several excitation sources were investigated including multiple impact, single impact, ambient traffic (from an adjacent bridge), test vehicle, and electro-dynamic shakers. While the authors noted that the modal frequencies and mode shapes extracted from the data of each test were consistent (i.e., not statistically different), significant changes were observed in the damping ratios which were correlated with excitation amplitude. 5.2.2 Frequency Band of Input Another source of epistemic uncertainty is related to the frequency band of the input, or more correctly, incorrect assumptions about the frequency band of the input. For example, during an ambient test of the Alamosa Canyon Bridge

using truck traffic on an adjacent bridge, Farrar et al. (2000) reported that some modes, identified through impact testing, could not be identified. The authors examined the PSD calculated from the ground motion (measured halfway between the two bridges) and noted relatively small amplitude at the particular frequency missed. As a result, it seems that either this frequency was not excited by the truck traffic on the adjacent bridge or this frequency was attenuated by the soil properties. In either case the assumption of broadband input was violated. It is important to point out that in many cases the data that allowed the authors to identify this gap would not be readily available (i.e., impact test results or suitable characterization of the input). 5.2.3 Spatial Characteristics of Input In addition to frequency and amplitude, significant epistemic uncertainty can stem from the spatial characteristics and directionality of the excitation. This is particularly important for ambient vibration testing of large bridges where traffic-induced excitation on the deck may overwhelm other ambient sources. A lack of transmissibility between the deck and the remaining structural sub-systems often limit the bandwidth and direction of excitation throughout a structure. In the case of forced vibration tests, shakers suitable for constructed facilities tend be very large, expensive, and time consuming to move and anchor to the structure. For example, during testing of the Seymour Bridge, the linear-mass shaker was located on the shoulder of the bridge so as to not impede the crawl-speed truck tests (Lenett, 1998). However, due to this location being at (or near) nodal points for several modes, only 4 of the first 13 modes of the structure were identified. Comparing the resulting modal flexibilities with static flexibilities (determined through crawl-speed truck tests), it was apparent that the modal flexibilities did not properly represent the structure due to modal truncation errors. Again, without an independent test to validate these results the user would be hard pressed to confidently and consistently identify such errors. 5.2.4 Time Requirement for Excitation A principle problem with any forced multiple input test is related to the time required to complete the test. As discussed previously, all structures in the field are non-stationary to a certain degree and if the test duration is such that appreciable changes in a structures properties occur, unreliable results can be expected. Lenett (1998) provides a very complete discussion of this issue related to the testing of the Seymour Bridge. During preliminary testing it was noted that this bridge had many closely spaced modes and thus a fine spatial resolution was selected for the impact testing. This resulted in 74 input locations which took over 35 hours to test completely (note that the Seymour Bridge is a relatively small bridge with a total length of 130 feet). Due to the non-stationary nature of the bridge, during this testing period the resulting modal flexibilities did not correlate well with the static flexibilities measured during a crawl-speed truck test. As a result, the investigators developed a more rapid test method that drastically reduced the time of testing and thus minimized the issues associated with non-stationarity due to ambient environmental conditions. Again, only through an independent check of the results (modal flexibility versus static flexibility) was this significant error identified

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS

409

5.3. Sensors, Data Acquisition, and Cabling In contrast to the previously discussed issues related to structural complexity and excitation, epistemic uncertainty associated with sensors, data acquisition and cabling plagues applications of St-ID for both constructed and manufactured systems. The following sections present a brief discussion of some sources of epistemic uncertainty associated with these systems. 5.3.1 Sensors Perhaps the most recognized source of epistemic uncertainty is associated with sensitivity bias errors resulting from improper sensor calibration. While standard procedures exist for characterizing and minimizing such uncertainty, it is important to point out that these calibration procedures are often carried out in environments very different from the application environment. In addition, due to the severe time limitations, access restrictions, and budget constraints typical of field testing investigations, several other sources of epistemic uncertainty exist. For example, in many cases these constraints only allow for so-called sensor-lite investigations, which can lead to spatial aliasing problems. In addition, sensor placement is often limited to convenient locations, which can lead to incomplete characterization if unmeasured portions of the structure dominate in modes captured from the accessible portion of the structure. As with all sources of epistemic uncertainty, it is often very difficult to identify the significance of the resulting errors associated with sensor issues. 5.3.2 Data Acquisition A typical data acquisition process consists of signal amplification, anti-alias filtering, analog-to-digital conversion, and time-windowing. Although some of these steps are used to eliminate or alleviate noises in measurements, they also introduce other types of errors in the measurements. For instance, the application of anti-alias filtering removes frequency contents above the Nyquist frequency (half of the sampling frequency) in order to prevent these higher frequencies from appearing as aliased lower frequency components. Because the anti-alias filtering has to be done before digitization, analog RC circuits are often used to implement the anti-alias filtering. The circuits are often modeled as second-order dynamic systems and they introduce amplitude and phase distortion of signals. While many researchers characterize data acquisition errors as noise, in fact most data acquisition problems in the field are due to spurious data spikes and energy at select harmonics that infiltrate the data (Zhang and Aktan, 2005). Such spurious signals should be considered a form of epistemic uncertainty. 5.3.3 Cabling The widespread use of electronic devices for measurements, communication, and computing results in interaction between these various systems that can adversely affect performance. Cables are particularly susceptible to electromagnetic interference because they are the longest parts of a system and act as antennas, picking up and radiating noise. Typically, three types of coupling (crosstalk) are considered for cabling. The first is capacitive or electric coupling, which results from the interaction of electric fields between cables.

The second is inductive or magnetic coupling, which is produced by the interaction between the magnetic fields of two cables. The last is a combination of electric and magnetic fields and is often called electromagnetic coupling or radiation. Depending on the type of error sources, they can produce either epistemic or aleatory uncertainties. They can often be minimized by proper grounding or shielding, but never removed completely. 6. Discussion While several examples of influential studies focused on the St-ID of constructed systems exist, the envisioned widespread use has not been realized as the development of St-ID techniques for constructed systems has lagged behind those associated with manufactured systems. Certainly the significant sources of uncertainty identified in the previous section have contributed to this; however, the fact that these uncertainties continue to persist more than 30 years after the first applications is indicative of a more fundamental problem. Specifically, that many researchers have discounted the value of conducting investigations into the behavior of actual constructed systems, and thus overcoming the identified uncertainties has not received the attention that it deserves. This position, often supported by pointing to the difficulty associated with acquiring reliable data in the field, has served as a self-fulfilling prophecy insuring that the impact of field research, taken as a whole, falls well short of what the level of investment would suggest. The result has been a fragmented and cautious use of field research, and a recasting of such studies as a means to simply gain detailed information about isolated structures. In this context, researchers have been able to claim success without facing the critically important and fundamental question of how can the results obtained from a specific structure be generalized to provide information applicable to other structures? In fact, many reject the possibility that results from field tests can be broadly applicable even though they provide the only objective link to actual constructed systems. In the case of laboratory experiments, the important question of generalization is addressed by claims that specimens and testing configurations represent some type of nominal conditions. What is often overlooked by these studies however, is that laboratory experiments tend to focus on fragmented portions of structures whose similitude (both in terms of behavior and levels of uncertainty) has not be established. Even for the exceedingly rare cases where complete full-scale structures have been tested in the laboratory, the exclusion of soil-structure interaction issues renders these studies unable to provide information on boundary conditions and global response quantities. This argument is not meant to imply that these tests are not valuable, but simply to point out their inherent limitations. In short, laboratory experiments must simply demonstrate a generalized specimen and testing configuration without rigorously illustrating both the similitude of behavior and the similitude of uncertainty to be viewed as broadly applicable. However, field experiments, for which similitude is guaranteed, are largely discounted in terms of their wider applicability because they are conducted on specific structures.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

410

The Shock and Vibration Digest / September 2006 Agbabian, M. S., Masri, S. F., Miller, R. K. and Caugher, T. K., 1990, System Identification Approach to Detection of Structural Changes, ASCE Journal of Engineering Mechanics, Vol. 117, 370 390. Aktan, A. E., 1995, Issues in Instrumented Health Monitoring, In Proceedings of the IABSE Symposium, San Francisco, on Extending the Life-Span of Structures, 911916. Aktan, A. E., Chuntavan, C., Toksoy, T. and Lee, K. L., 1993a, Structural Identification of a Steel-Stringer Bridge for Nondestructive Evaluation, Transportation Research Record (TRB), Issue 1393, National Academy Press, Washington, D.C., 175185. Aktan, A. E., Zwick, M. J., Miller, R. A. and Shahrooz, B. M., 1993b, Nondestructive and Destructive Testing of a Decommissioned RC Slab Highway Bridge and Associated Analytical Studies, Paper presented at TRB 92 and published in Transportation Research Record (TRB), Issue 1371, National Academy Press, Washington, D.C., 142153. Aktan, A. E., Farhey, D. N. and Dalal, V., 1995a, Issues in Rating SteelStringer Bridges, Transportation Research Record (TRB), Issue 1476, National Academy Press, Washington, D.C., 129138. Aktan, A. E., Lee, K. L., Naghavi, R. and Hebbar, K., 1995b, Nondestructive and Destructive Experimental Studies of Two 80-Year Old Bridges, Transportation Research Record (TRB), Issue 1460, National Academy Press, Washington D.C., 6272. Aktan, A. E., Naghavi, R., Farhey, D. N., Lee, K. L., Aksel, T. and Hebbar, K., 1995c, Nondestructive/ Destructive Tests and Associated Studies on Two Decommissioned Steel Truss Bridges, Final Report, FHWA/OH-95/013. Aktan, A. E., Farhey, D. N., Brown, D. L., Dalal, V., Helmicki, A. J., Hunt, V. J. and Shelley, S. J., 1996, Condition Assessment for Bridge Management, ASCE Journal of Infrastructure Systems, Vol. 2(3), 108117. Aktan, A. E., Farhey, A. N., Helmicki, A. J., Brown, D. L., Hunt, V. J., Lee, K. L. and Levi, A., 1997a, Structural Identification for Condition Assessment: Experimental Arts, ASCE Journal of Structural Engineering, Vol. 123(12), 16741684. Aktan, A. E., Helmicki, A. J., Hunt, V. J., Catbas, N., Lenett, M. and Levi, A., 1997b, Structural Identification for Condition Assessment of Civil Infrastructure, In Proceedings of the American Control Conference, Albuquerque, NM. Aktan, A. E., Catbas, N., Turer, A. and Zhang, Z., 1998a, Structural Identification: Analytical Aspects, ASCE Journal of Structural Engineering, Vol. 124(7), 817829. Aktan, A. E., Helmicki, A. J. and Hunt, V. J., 1998b, Issues in HealthMonitoring for Intelligent Infrastructure, Journal of Smart Materials and Structures, Vol. 8(5), 674692. Aktan, A. E., Catbas, F. N., Grimmelsman, K. A. and Tsikos, C. J., 2000b, Issues in Infrastructure Health Monitoring for Management, Journal of Engineering Mechanics, Vol. 126(7), 711724. Aktan, A. E., Grimmelsman, K. A., Barrish, R. A., Catbas, F. N. and Tsikos, C. J., 2000a, Structural Identification of a Long-Span Truss Bridge, Transportation Research Record (TRB), Issue 1696, National Academy Press, Washington, D.C., 210218. Aktan, A. E., Catbas F. N., Grimmelsman, K. A. and Pervizpour, M., 2002, A Model Health Monitoring Guide for Major Bridges, Report to FWHA, DOT/FHWA Solicitation: DTFH6101-Q-00072, www.di3. drexel.edu. Aktan, A. E., Ellingwood, B. and Kehoe, B., 2006, Performance-Based Engineering of Constructed Systems, Journal of Structural Engineering, Committee Report, (in press). Alampalli, S., 1998, Influence of In-Service Environment on Modal Parameters, In Proceedings of the 16th International Modal Analysis Conference (IMAC 16), Santa Barbara, CA, 11116. Al-Khalidy, A., Noori, M., Hou, Z., Yamamoto, S., Masuda, A. and Sone, A., 1997, Health Monitoring Systems of Linear Structures Using Wavelet Analysis, In Proceedings of the International Workshop on Structural Health Monitoring, Stanford University, CA,164175 Al-Mahaidi, R., Taplin, G. and Giufre, A., 2000, Load Distribution and Shear Strength Evaluation of an Old Concrete T-Beam Bridge, Transportation Research Record (TRB), Issue 1696, 5261. Andersen, P. and Kirkegaard, P. H., 1998, Statistical Damage Detection of Civil Engineering Structures Using ARMAV Models, In Proceedings of the 16th International Modal Analysis Conference (IMAC 16), Santa Barbara, CA, 356362. Ang, A. H.-S. and De Leon, D., 2005, Modeling and Analysis of Uncertainties for Risk-Informed Decisions in Infrastructures Engineering, Structure and Infrastructure Engineering, Vol. 1(1), 1931. Asmussen, J. C., Brincker, R. and Rytter, A., 1998a, Ambient Model Testing of the Vestvej Bridge Using Random Decrement, In Pro-

Although this type of double standard may seem inexplicable, it is precisely the difficulty in establishing the reliability of field data and overcoming the identified uncertainties that furthers this position. Consider that in cases where similar structures are tested and compared or where modelbased simulations are used, the testing results rarely display strong correlations, which leads the community to either conclude significant errors have occurred or that the test structure is not representative of nominal conditions. In reality however, the authors believe that these differences are not representative of anomalies, but rather realistic and important behavior mechanisms. As a result, they should not be discounted as overly specific, but they should be embraced and examined in the hopes of shedding light on actual structural behavior. If civil engineers are ever to shed the undesirable distinction as the oldest engineering discipline, with the least understanding of the actual mechanical characteristics, behaviors and performances of the systems they design, St-ID of actual constructed systems must become commonplace. In recognition of the importance of furthering this view, which was also recently articulated by the ASCE Performance-Based Design Committee (Aktan, et al, 2006), the ASCE-SEI has recently re-established the St-ID of Constructed Facilities Committee. From 1997-2002, the elder writer founded and co-chaired this committee together with Dr. J. Yao. This committee developed a draft state-of-the-art report but became inactive in 2002 and could not disseminate the report. The primary goal of the re-established committee is to invite engineers from all engineering subdisciplines to develop a consensus terminology in conjunction with minimum quality standards for integrating the entire St-ID cycle (Figure 2). The authors would like to invite colleagues from all disciplines interested in further the art of St-ID of constructed facilities to consider joining this committee. Acknowledgements The authors are grateful to Dr Inman for inviting this paper, and to Drs de Roeck, Brownjohn, Fujino and Wenzel for their collaboration and guidance in the writing of this paper. The authors are grateful to the FHWA, NSF and various state and local infrastructure agencies that have supported their research on structural identification of constructed systems.

References
Abe, M., Fujino, Y., Yanagihara, M. and Sato, M., 2000, Monitoring of Hakucho Suspension Bridge by Ambient Vibration Measurement, In Proceedings of SPIE The International Society for Optical Engineering, Vol. 3995, 237244. Abdel-Ghaffar, A. M. and Housner, G. W., 1978, Ambient Vibration Tests of Suspension Bridge, ASCE Journal of the Engineering Mechanics Division, Vol. 104, 983999. Abdel-Ghaffar, A. M. and Scanlan, R. H., 1985a, Ambient Vibration Studies of Golden Gate Bridge: I. Suspended Structure, ASCE Journal of Engineering Mechanics, Vol. 111, 463482. Abdel-Ghaffar, A. M. and Scanlan, R. H., 1985b, Ambient Vibration Studies of Golden Gate Bridge: II. Pier-Tower Structure, ASCE Journal of Engineering Mechanics, Vol. 111, 483499. Abdel Wahab, M. M. and De Roeck, G., 1999, Damage Detection in Bridges Using Modal Curvatures: Application to a Real Damage Scenario, Journal of Sound and Vibration, Vol. 226(2), 217235.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS ceedings of the 16th International Modal Analysis Conference (IMAC 16), Santa Barbara, CA, 922928. Asmussen, J. C., Ibrahim, S. R. and Brincker, R., 1998b, Random Decrement: Identification of Structures Subjected to Ambient Excitation, In Proceedings of the 16th International Modal Analysis Conference (IMAC 16), Santa Barbara, CA, 914921. Banish, R. A., Jr., Grimmelsman, K. A. and Aktan, A. E., 2000, Instrumented Monitoring of the Commodore Barry Bridge, In Proceedings of the Fifth International Symposium on Nondestructive Evaluation and Health Monitoring of Aging Infrastructure, The International Society for Optical Engineering, (SPIE), Newport Beach, CA, 112126. Beck, J. L., 1990, Statistical System Identification of Structures, In Proceedings of ICOSSAR 89, the 5th International Conference on Structural Safety and Reliability, San Fransisco, CA, 13951402. Beck, J. L. and Katafygiotis, L. S., 1998, Updating Models and Their Uncertainties: Bayesian Statistical Framework, Journal of Engineering Mechanics, Vol. 124(4), 455461. Bernard, K. J., Culmo, M. P. and DeWolf, J. T., 1997, Strain Monitoring to Evaluate Steel Bridge Connections, In Building to Last: the Proth ceedings of the 15 ASCE Structures Congress, Portland, OR, 919 923. Bertero, V. V., Aktan, A. E., Charney, F. A. and Sause, R., 1985, Earthquake Simulator Tests and Associated Experimental, Analytical and Correlation Studies of a 1/5 Scale Replica Model of a Full Scale, 7Story R/C Frame-Wall Structure, In J. K. Wight (Ed.) Earthquake Effects on Reinforced Concrete Structures: U.S.-Japan Research, American Concrete Institute. Billing, J. R., 1984, Dynamic Loading and Testing of Bridges in Ontario, Canadian Journal of Civil Engineering, Vol. 11, 833843. Billington, D. P., 1983, The Tower and the Bridge, Basic Books, New York. Billington, D. P., 1979, Robert Maillarts Bridges: The Art of Engineering, Princeton University Press, Princeton, NJ. Biswas, M., Pandey, A. K. and Samman, M. M., 1989, Diagnostic Experimental Spectral/Modal Analysis of a Highway Bridge, International Journal of Analytical and Experimental Modal Analysis, Vol. 5(1), 3342. Bodeux, J. B. and Golinval, J. C., 2001, Application of ARMAV Models to Identification and Damage Detection of Mechanical and Civil Engineering Structures, Smart Materials and Structures, Vol. 10, 479489. Bollo, M. E., Mahin, S. A., Moehle, J. P., Stephen, R. M. and Qi, X., 1990, Observations and Implications of Tests on the Cypress Street Viaduct Test Structure, UCB/EERC-90/21. Bolton, R., Stubbs, N. and Sikorsky, C., 2002, Automation of Modal Property Extraction for a Permanently Instrumented Highway Bridge, In Proceedings of the 20th International Modal Analysis Conference (IMAC 20), Los Angeles, CA, 12171223. Brincker, R., Frandsen, J. B. and Andersen, P., 2000, Ambient Response Analysis of the Great Belt Bridge, In Proceedings of the 18th International Modal Analysis Conference (IMAC 18), San Antonio, TX, 2632. Brincker, R., Ventura, C. E. and Andersen, P., 2001, Damping Estimation th by Frequency Domain Decomposition, In Proceedings of the 19 International Modal Analysis Conference (IMAC 19), Kissimmee, FL, 698703. Brownjohn, J. M. W., Chandu, R., Dumanoglu, A. A. and Taylor, C. A., 1986, Ambient Vibration Testing on the Humber Suspension Bridge, In Proceedings of the 8th European Conference on Earthquake Engineering, Laborotrio Nacional de Engenharia Civil, Lisbon, Portugal, Vol.4, 8996. Brownjohn, J. M. W., Dumanoglu, A. A., Severn, R. T. and Taylor, C. A., 1987, Ambient Vibration Measurements of the Humber Suspension Bridge and Comparison with Calculated Results, Proceedings of the Institution of Civil Engineers, Vol. 83(2), 561600. Brownjohn, J. M. W., Dumanoglu, A. A., Severn, R. T. and Blakeborough, A., 1989, Ambient Vibration Survey of the Bosporus Suspension Bridge, Earthquake Engineering and Structural Dynamics, Vol. 18, 263283. Brownjohn, J. M. W., Dumanoglu, A. A. and Severn, R. T., 1992a, Ambient Vibration Survey of the Fatih Sultan Mehmet (Second Bosporus) Suspension Bridge, Earthquake Engineering and Structural Dynamics, Vol. 21, 907924. Brownjohn, J. M. W., Severn, R. T. and Dumanoglu, A. A., 1992b, FullScale Dynamic Testing of the Second Bosporus Suspension Bridge, In Proceedings of the 10th World Conference on Earthquake Engineering, Vol.5, 26952700.

411

Brownjohn, J. M. W., Bocciolone, M., Curami, A., Falco, M. and Zasso, A., 1994, Humber Bridge Full-Scale Measurement Campaigns 19901991, Journal of Wind Engineering and Industrial Aerodynamics, Vol. 52(13), 185218. Brownjohn, J. M. W., Lee, J. and Cheong, B., 1999, Dynamic Performance of a Curved Cable-Stayed Bridge, Engineering Structures, Vol. 21(11), 10151027. Brownjohn, J. M. W., Moyo, P., Lu, Y. and Tan, B. L., 2002, DynamicsBased Condition Assessment of a Highway Bridge to Assess the Effect of Upgrading, Proceedings of SPIE The International Society for Optical Engineering, Vol. 4753(1), 644649. Brownjohn, J. M. W., Moyo, P., Omenzetter, P. and Lu, Y., 2003, Assessment of Highway Bridge Upgrading by Dynamic Testing and FiniteElement Model Updating, Journal of Bridge Engineering, Vol. 8(3), 162172. Bucher, C., Huth, O. and Macke, M., 2003, Accuracy of System Identification in the Presence of Random Fields, In Proceedings of the 9th International Conference on the Applications of Statistics and Probability (ICASP 9). Buckland, P. G., Hooley, R., Morgenstern, B. D., Rainer, J. H. and Van Selst, A. M., 1979, Suspension Bridge Vibrations: Computed and Measured, ASCE Journal of the Structural Division, Vol. 105, 859874. Buckle, I. G., Douglas, B. M., Saiidi, M., Richardson, J. A. and Butterworth J. W., 1986, Field Tests of a Curved Box Girder Bridge Using Simulated Earthquake Loads, In Proceedings of the 8th European Conference on Earthquake Engineering: Laboratori Nacional de Engenharia Civil, Lisbon, Portugal, Vol. 4, 18. Burdette, E. G. and Goodpasture, D. W., 1972, Comparison of Measured and Computed Ultimate Strengths of Four Highway Bridges, Highway Research Record, Issue 382, 3849. Cantieni, R., 1984, Dynamic load testing of highway bridges, Transportation Research Record (TRB), Issue 950, 141148. Cantieni, R. and Pietrzko S., 1993, Modal Testing of a Wooden Footbridge Using Random Excitation, In Proceedings of the 11th International Modal Analysis Conference (IMAC 11), Kissimmee, FL. Catbas, F. N. et al., 1997, Modal Analysis of Multireference Impact Test Data for Steel Stringer Bridges, In Proceedings of the 15th International Modal Analysis Conference (IMAC 15), Orlando, FL, 381389. Catbas, F. N., Lenett, M., Aktan, A. E., Brown, D. L., Helmicki, A. J. and Hunt, V. J., 1998, Damage Detection and Condition Assessment of Seymour Bridge, In Proceedings of the 16th International Modal Analysis Conference (IMAC 16), 16941702. Catbas, F. N., Grimmelsman, K. A. and Aktan, A. E., 2000, Structural Identification of the Commodore Barry Bridge, Keynote Address, International Conference on Advances in Structural Dynamics, Hong Kong, December 1315. Catbas, F. N. and Aktan, A. E., 2002, Condition and Damage Assessment: Issues and Some Promising Indices, ASCE Journal of Structural Engineering, Vol. 128(8). Catbas, F. N., Ciloglu, S. K., Grimmelsman, K., Pan, Q., Pervizpour, M. and Aktan, A. E., 2003, Limitations in the Structural Identification of Long-Span Bridges, In Proceedings of the International Workshop on Structural Health Monitoring of Bridges and Colloquium on Bridge Vibration, Kitami, Japan. Catbas, F. N., Brown, D. L. and Aktan, A. E., 2004, Parameter Estimation for Multiple-Input Multiple-Output Modal Analysis of Large Structures, ASCE Journal of Engineering Mechanics, Vol. 130(8), 921930. Chang, C. C., Chang, T. Y. P. and Zhang, Q. W., 2001, Ambient Vibration of Long-Span Cable-Stayed Bridge, ASCE Journal of Bridge Engineering, Vol. 6(1), 4653. Chassiakos, A. G, Masri, S. F., Smyth, A. and Anderson, J. C., 1995, Adaptive Methods for Identification of Hysteretic Structures, In Proceedings of the American Control Conference (ACC95), Seattle, WA. Chen, J., Xu, Y. L. and Zhang, R. C., 2004, Modal Parameter Identification of Tsing Ma Suspension Bridge under Typhoon Victor: EMDHT Method, Journal of Wind Engineering and Industrial Aerodynamics, Vol. 92(10), 805827. Christensen, C., Halling, M. W. and Womack, K. C., 2001, Dynamic Properties of a Full-Scale Bridge Using Forced Vibration Testing, In Proceedings of the 19th International Modal Analysis Conference (IMAC 19), Kissimmee, FL. Cornwell, P., Farrar, C. R., Doebling, S. W. and Sohn, H., 1999, Environmental Variability of Modal Properties, Experimental Techniques, Vol. 23(6), 4548. Crouse, C. B., Hushmand, B. and Martin, G.R., 1987, Dynamic SoilStructure Interaction of a Single-Span Bridge, Earthquake Engineering and Structural Dynamics, Vol. 15, 711729.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

412

The Shock and Vibration Digest / September 2006 Filiatrault, A., Tinawi, R., Felber, A., Ventura, C. E. and Stiemer, S.F., 1993a, Modal Analysis and Testing of the Cable-Stayed Shipshaw Bridge in Jonquiere, Quebec, In Proceedings of the 11th International Modal Analysis Conference (IMAC 11), Kissimmee, FL, 1420. Filiatrault, A., Tinawi, R., Felber, A., Ventura, C. E. and Stiemer, S. F., 1993b, Modal Analysis and Testing of the Cable-Stayed Shipshaw Bridge in Jonquiere, Quebec, In Proceedings of the 11th Internation Modal Analysis Conference (IMAC), Kissimmee, FL, 1420. Foutch, D. A., Luco, J. E., Trifunac, M. D. and Udwadia, F. E., 1975, Full-Scale, Three-Dimensional Tests of Structural Deformations During Forced Excitation of a Nine-Story Reinforced Concrete Building, In Proceedings of the US National Conference on Earthquake Engineering, Ann Arbor, MI, 206215. Foutch, D. A., Hjelmstad, K. D., Del Valle Caldern, E., Figueroa Gutirrez, E. and Downs, R. E., 1989, The Mexico Earthquake of September 19, 1985Case Studies of Seismic Strengthening for Two Buildings in Mexico City, Earthquake Spectra, Vol. 5(1), 153174. Fu, Y. and DeWolf, J. T., 2001, Monitoring and Analysis of a Bridge with Partially Restrained Bearings, ASCE Journal of Bridge Engineering, Vol. 6(1). Fujino, Y. and Abe, M., 2002, Vibration-Based Structural Health Monitoring of Civil Infrastructures, In Proceedings of Structural Health Monitoring Workshop, Winnipeg, Canada, 4555. Fujino, Y., Nakamura, S. I., Shibuya, H., Sato, M., Yanagihara, M. and Sakamoto, Y., 1999, Forced and Ambient Vibration Tests of Hakucho Suspension Bridge, In Structural Engineering in the 21st Centh tury: the Proceedings of the 17 ASCE Structures Congress, New Orleans, LA, 328331. Fujino, Y., Abe, M., Shibuya, H., Yanagihara, M., Sato, M., Nakamura, S. and Sakamoto, Y., 2000a, Forced and Ambient Vibration Tests and Vibration Monitoring of Hakucho Suspension Bridge, Transportation Research Record (TRB), Issue 1696, 5763. Fujino, Y., Murata, M., Okano, S. and Takeguchi, M., 2000b, Monitoring System of the Akashi Kaikyo Bridge and Displacement Measurement Using GPS, In Proceedings of SPIE The International Society for Optical Engineering, Vol. 3995, 229236. Galambos, T. V. and Mayes, R. L., 1978, Dynamic Tests of a Reinforced Concrete Building, School of Engineering and Applied Science, Department of Civil Engineering, Washington University, Research Report No. 51. Gates, J. H. and Smith, M. J., 1982, Verification of Dynamic Modeling Methods by Prototype Experiments, FHWA/CA/SD-82/07. Gilani, A. S., Mahin, S. A., Fenves, G. L., Aiken, I. D. and Chavez, J. W., 1995, Field Testing of Bridge Design and Retrofit Concepts: Part 1 of 2: Field Testing and Dynamic Analysis of a Four-Span Seismically Isolated Viaduct in Walnut Creek, California, UCB/EERC-95/14. Green, M.F. and Cebon, D., 1993, Modal Testing of Two Highway bridges, In Proceedings of the 11th International Modal Analysis Conference on Structural Dynamics (IMAC 11), Kissimmee, FL, 838844. Grimmelsman, K. A. and Aktan, A. E., 2005, Impacts and Mitigation of Uncertainty for Improving the Reliability of Field Measurements, In nd Proceedings of the 2 International Conference on Structural Health Monitoring of Intelligent Infrastructure (SHMII 2), Shenzhen, China. Gundy, W. E., Howard, G. E., Ibanez, P., Keowen, R. S. and Smith, C. B., 1981, High Level Seismic Test Using Shaker and Snapback Methods on a Hot Plant, In Transactions of the International Conference on Structural Mechanics in Reactor Technology. Volume K: Seismic Response Analysis of Nuclear Power Plant Systems, Paris, France, p. 8. Gurley, K. and Kareem, A., 1999, Applications of Wavelet Transform in Earthquake, Wind and Ocean Engineering, Engineering Structures, Vol. 21(2), 149167. Hales, M. B. and Halling, M. W., 2002, Long-Term Bridge Vibration Monitoring Using Strong Motion Sensors, In Proceedings of the 20th International Modal Analysis Conference on Structural Dynamics (IMAC 20), Los Angeles, CA. Halling, M. W., Muhammad, I. and Womack, K. C., 2001, Dynamic Field Testing for Condition Assessment of Bridge Bents, Journal of Structural Engineering, Vol. 127(2), 161167. Halling, M. W., Ball, A. W., Esplin, R. B., Petty, T. S., Troy, D. and Hales, M. W., 2004, Dynamic Health Monitoring and Modeling of a FullScale Bridge, In Building on the Past: Securing the Future: Proceedings of the 2004 ASCE Structures Congress, Nashville, TN, 12591265. Hart, G. C. and Yao, J. T. P., 1977, System Identification in Structural Dynamics, ASCE Journal of the Engineering Mechanics Division, Vol. 103(EM6), 10891104.

Culmo, M. P., DeWolf, J.T. and DelGrego, M. R., 2004, Behavior of Steel Bridges Under Superload Permit Vehicles, Transportation Research Record (TRB), Issue 1892, 107114. De Roeck, G. 2003, The State-of-the-Art of Damage Detection by Vibration Monitoring: the SIMCES Experience, Journal of Structural Control, Vol. 10, 127143. De Roeck, G., Peeters, B. and Maeck, J., 2000, Dynamic Monitoring of Civil Engineering Structures, Computational Methods for Shells and Spatial Structures, IASS-IACM. DeWolf, J. T., Lindsay, T.R., and Culmo, M.P., 1997, Fatigue Evaluations in Steel Bridges Using Field Monitoring Equipment, In Building to th Last: the Proceedings of the 15 ASCE Structures Congress, Portland, OR, 2630. Doebling, S. W., Farrar, C.R. and Goodman, R., 1997, Effects of Measurement Statistics on the Detection of Damage in the Alamosa Canyon Bridge, In Proceedings of the 15th International Modal Analysis Conference (IMAC 15), Orlando, FL. Doebling, S. W., Farrar, C. R. and Prime, M. B., 1998, A Summary Review of Vibration-Based Damage Identification methods, The Shock and Vibration Digest, Vol. 30(2), 91105. Doebling, S. W., Farrar, C. R., Aktan, A. E., Beck, J., Cornwell, P., Helmicki, A., Safak, E. and Yao, J., 2000, The State of the Art in Structural Identification of Constructed Facilities, A Draft Report by the ASCE Committee on Structural Identification of Constructed Facilities. Douglas, B. M., 1976, Quick Release Pullback Testing and Analytical Seismic Analysis of Six Span Composite Girder Bridge, FHWARD16-173. Douglas, B. M. and Reid, W. H., 1982, Dynamic Testing and System Identification of Bridges, ASCE Journal of Structural Engineering, Vol. 108, 22952312. Douglas, B. M., Brown, C. D. and Gordon, M. L., 1981, Experimental Dynamics of Highway Bridges, In Proceedings of the Conference on Dynamic Response of Structures: Experimentation, Observation, Prediction and Control, 698712. Eberhard, M. O. and Marsh, M. L., 1997a, Transverse-Load Response of a Reinforced Concrete Bridge, ASCE Journal of Structural Engineering, Vol. 123(4), 451460. Eberhard, M. O. and Marsh, M. L., 1997b, Transverse-Load Response of Two Reinforced Concrete Bents, ASCE Journal of Structural Engineering, Vol. 123(4), 461468. Eves, H., 1990, An Introduction to the History of Mathematics with Cultural Connections (Sixth edition), Saunders College Publishing, Fort Worth, TX. Farhey, D., Thakur, A., Buchanan, R., Aktan, A. E. and Jayaraman, N., 1997, Deterioration Assessment for Steel Bridges, ASCE Journal of Bridge Engineering, Vol. 2(3), 116124. Farhey, D. N., Naghavi, R., Levi, A., Thakur, A. M., Pickett, M. A., Nims, D. K. and Aktan, A. E., 2000, Deterioration Assessment and Rehabilitation Design of an Existing Steel Bridge, Journal of Bridge Engineering, Vol. 5(1), 3948. Farrar, C. R. and Doebling, S. W., 1998, Damage Detection II: Field Applications to Large Structures, In Proceedings of the NATO Advanced Study Institute on Modal Analysis and Testing. Farrar, C. R. and Jauregui, D. A., 1996, Damage Detection Algorithms Applied to Experimental and Numerical Modal Data from the I-40 Bridge, Los Alamos National Laboratory Report LA-13074-MS. Farrar, C. R. and Jauregui, D.A., 1998a, Comparative Study of Damage Identification Algorithms Applied to a Bridge: II. Numerical Study, Smart Materials and Structures, Vol. 7, 720731. Farrar, C. R. and Jauregui, D.A., 1998b, Comparative Study of Damage Identification Algorithms Applied to a Bridge: I. Experiment, Smart Materials and Structures, Vol. 7, 704719. Farrar, C. R., Baker, W. E., Bell, T. M., Cone, K. M., Darling, T. W., Duffey, T. A., Eklund, A. and Migliori, A., 1994, Dynamic Characterization and Damage Detection in the I-40 Bridge over the Rio Grande, Los Alamos National Laboratory Report LA-12767-MS. Farrar, C. R., Duffey, T. A., Goldman, P. A., Jauregui, D.A. and Vigil, J.S., 1996, Finite element analysis of the I-40 Bridge over the Rio Grande, Los Alamos National Laboratory Report LA-12979-MS. Farrar, C. R., Doebling, S. W., Cornwell, P.J., and Straser, E.G., 1997, Variability of modal parameters measured on the Alamosa Canyon Bridge, In Proceedings of the 15th International Modal Analysis Conference (IMAC 15), Orlando, FL. Farrar, C. R., Cornwell, P. J., Doebling, S. W. and Prime, M. B., 2000, Structural Health Monitoring Studies of the Alamosa Canyon and I40 Bridges, Los Alamos National Laboratory Report LA-13635-MS.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS Hornbuckle, J., Mercurio T., Howard, G, Ibanez, P., Smith, C. B. and Vasudevan, R., 1973, Forced Vibration Tests and Analyses of Nuclear Power Plants, Nuclear Engineering and Design, Vol. 25(1), 51 93. Hou, Z. K., Noori, M. and St. Amand, R., 2000, Wavelet-Based Approach for Structural Damage Detection, ASCE Journal of Engineering Mechanics, Vol. 126(7), 677683. Huang, N. E., Shen, Z., Long, S. R., Wu, M. C., Shih, H. H., Zheng, Q., Yen, N.-C., Tung, C. C. and Liu, H. H., 1998, The Empirical Mode Decomposition and the Hilbert Spectrum for Non-Linear and NonStationary Time Series Analysis, Proceedings of the Royal Society, Vol. 454, 903995. Hudson, D. E., 1970, Dynamic Tests of Full-Scale Structures, In R. L. Wiegel (Ed.), Earthquake Engineering, Prentice-Hall, Englewood Cliffs, NJ. Hudson, D. E., 1977, Dynamic Tests of Full-Scale Structures, ASCE Journal of the Engineering Mechanics Division, Vol. 103, 11411157. Ibanez, P., Chien, S., Dinyavari, M., Dobbs, M., Gundy, W., Howard, G., Keownen, R., Rentz, P., Smith, C., Stoessel, J., Walton, W. and Sires-Yifat, C., 1981, Methods and Benefits of Experimental Structural Dynamic Evaluation of Nuclear Power Plants, Nuclear Engineering and Design, Vol. 64(1), 132 Iwasaki, T., Penzien, J. and Clough, R., 1972, Literature Survey Seismic Effects on Highway Bridges, EERC report 7211, University of California, Berkeley. Jayakumar, P. and Beck, J. L., 1988, System Identification Using Nonlinear Structural Models, In H.G. Natke and J.T.P. Yao (Eds.), Structural Safety Evaluation Based on System Identification Approaches, Vieweg-Verlag, Wiesbaden, Germany. Jorgenson, L. J. and Larson, W., 1972, Field Testing of a Reinforced Concrete Bridge to Collapse, Transportation Research Record (TRB), 607, 6671. Kapania, R. K. and Park, S., 1997, Parametric Identification of Nonlinear Structural Dynamic Systems Using Time Finite Element Method, American Institute of Aeronautics and Astronautics, Vol. 35(4). Kim, B. H., Stubbs, N. and Park, T., 2005, A New Method to Extract Modal Parameters Using Output-Only Responses, Journal of Sound and Vibration, Vol. 282, 215230 Ko, J. M., Ni, Y. Q. and Chan, T. H. T., 2000, Feasibility of Damage Detection of Tsing Ma Bridge Using Vibration Measurements, In Proceedings Of SPIE Nondestructive Evaluation of Highways, Utilities and Pipelines IV, Newport Beach, CA, Vol. 3995, 370381. Kohoutek, R., 1993, Tests on Bridge over Talbragar River at Dubbo, In Proceedings of the 11th International Modal Analysis Conference on Structural Dynamics (IMAC 11), Kissimmee, FL, 11681174. Kossiakoff, A. and Sweet, W.N., 2002, Systems Engineering Principles and Practice, Wiley. Kou, J. W. and DeWolf, J. T., 1997, Vibrational Behavior of Continuous Span Highway Bridge Influencing Variables, ASCE Journal of Structural Engineering, Vol. 123(3), 333344. Kumarasena, T., Scanlan, R.H. and Morris, G.R., 1989, Deer Isle Bridge: Field Computed Vibrations, ASCE Journal of Structural Engineering, Vol. 115, 23132328. Kumarasena, T., Scanlan, R. H., and Ehsan, F., 1991, Wind-Induced Motions of Deer Isle Bridge, ASCE Journal of Structural Engineering, Vol. 117(11), 33563374. Kuribayashi, E., and Iwasaki, T., 1973, Dynamic Properties of Highway Bridges, In Proceedings of the 5th World Conference on Earthquake Engineering, Rome, Italy, 938941. Kwong, H. S., Lau, C. K. and Wong, K. Y., 1995, Monitoring System for Tsing Ma Bridge, In Restructuring: America and Beyond: Proceedings of the 13th Structures Congress, Boston, MA, Vol.1, 264267. Lauzon, R. G., and DeWolf, J. T., 1995, Nondestructive Evaluation with Vibrational Analysis, In H. V. GangaRao (Ed.), Nondestructive Testing Methods for Civil Infrastructure, A Collection of Expanded Papers on Nondestructive Testing from the 11th ASCE Structures Congress 93, 1730. Lee, P. K., Ho, D. and Chung, H. W., 1987, Static and Dynamic Tests of Concrete Bridge, ASCE Journal of Structural Engineering, Vol. 113, 6173. Lenett, S. M., 1998, Global Condition Assessment Using Modal Analysis and Flexibility, PhD Thesis, University of Cincinnati, OH. Lenett, M., Catbas, N., Hunt, V., Aktan, A. E., Helmicki, A., and Brown, D., 1997, Issues in Multireference Impact Testing of Steel-Stringer Bridges, In Proceedings of the 15th International Modal Analysis Conference on Structural Dynamics (IMAC 15), Orlando, FL, 374 380.

413

Lenett, M. S., Griessman, A., Helmicki, A. J. and Aktan, A. E., 1999, Subjective and Objective Evaluations of Bridge Damage, Transportation Research Record (TRB), Issue 1688, 7686. Leonard, D. R. and Eyre, R., 1975, Dynamic Tests on Highway Bridges Test Procedures and Equipment, TRRL Laboratory Report 682, Transport and Road Research Laboratory, Berkshire, UK. Liu, S.C. and Yao, J.T.P. 1978, Structural Identification Concept, ASCE Journal of the Structural Division, Vol. 104(12), 18451858. Luco, J. E., Trifunac, M. D. and Wong, H. L., 1988, Isolation of SoilStructure Interaction Effects by Full-Scale Forced Vibration Tests, Earthquake Engineering & Structural Dynamics, Vol. 16(1). Maeck, J. and De Roeck, G., 2003, Damage Assessment Using Vibration Analysis on the Z24-Bridge, Mechanical Systems and Signal Processing, Vol. 17(1), 133142. Maeck, J., Peeters, B. and De Roeck, G., 2001, Damage Identification on the Z24 Bridge Using Vibration Monitoring, Smart Materials and Structures, Vol. 10, 512517. Mahmoud, M., Abe, M. and Fujino, Y., 2001, Analysis of Suspension Bridge by Ambient Vibration Measurement Using Time Domain Method and its Application to Health Monitoring, In Proceedings of the 19th International Modal Analysis Conference (IMAC 19), Kissimmee, FL, 504510. Masri, S. F., Nakamura, M., Chassiakos, A. G. and Caughey, T. K., 1996, Neural Network Approach to Detection of Changes in Structural Parameters, Journal of Engineering Mechanics, Vol. 122(4), 350 360. Masri, S. F., Sheng, L.-H., Caffrey, J. P., Nigbor, R. L., Wahbeh, M. and Abdel-Ghaffar, A. M., 2004, Application of a Web-Enabled RealTime Structural Health Monitoring System for Civil Infrastructure Systems, Smart Materials and Structures, Vol. 13(6), 12691283. Mayes, R. L., 1995, An Experimental Algorithm for Detecting Damage Applied to the I-40 Bridge over the Rio Grande, In Proceedings of the 13th International Modal Analysis Conference (IMAC 13), Nashvile, TN, 21925. Mayes, R. L. and Nusser, M. A., 1994, The Interstate-40 Bridge Shaker Project, Sandia National Laboratory Report SAND940228. McClure, R. M. and West, H. H., 1984, Full-Scale Test of a Prestressed Concrete Segmental Bridge, Canadian Journal of Civil Engineering, Vol. 11, 505515. McLamore, V. R., Hart, G. C. and Stubbs, I. R., 1971, Ambient Vibration of Two Suspension Bridges, ASCE Journal of the Structural Division, Vol. 97, 25672582. Miller, R. A., Aktan, A. E. and Sharooz, B. M., 1992, Nondestructive and Destructive Testing of a Three Span Skewed RC Slab Bridge, In Proceedings of the ASCE Conference on Nondestructive Testing of Concrete Elements and Structures, San Antonio, TX, 15016. Miller, R. A., Aktan, A. E. and Shahrooz, B., 1994, Destructive Testing of a Decommissioned Concrete Slab Bridge, ASCE Journal of Structural Engineering, Vol. 120(7). Moon, F. L., Yi, T., Leon, R. T. and Kahn, L. F., 2006, Recommendations for the Seismic Evaluation and Retrofit of Low-Rise URM Structures, ASCE Journal of Structural Engineering, Vol. 132(5). Moyo, P., Brownjohn, J. M. W. and Omenzetter, P., 2004, Highway Bridge Live Loading Assessment and Load Carrying Capacity Estimation Using a Health Monitoring System, Structural Engineering and Mechanics, Vol. 18(5), 609626. Muria-Vila, D., Gomez, R. and King, C., 1991, Dynamic Structural Properties of Cable-Stayed Tampico Bridge, ASCE Journal of Structural Engineering, Vol. 117, 33963416. Naghavi, R. and Aktan, A. E.. 2003, Nonlinear Behavior of Existing Heavy-Class Steel Truss Bridges, ASCE Journal of Structural Engineering, Vol. 129(8). Nakamura, M., Masri, S. F., Chassiakos, A. G. M. and Caughey, T. K., 1998, Method for Nonparametric Damage Detection Through the Use of Neural Networks, Earthquake Engineering and Structural Dynamics, Vol. 27, 9971010. Natke, H. G. and Yao, J. T. P., 1986, Research Topics in Structural Identification, Proceedings of the Third ASCE Conference on Dynamic Response of Structures, Los Angeles, CA, 542550. Natke, H. G. and Yao, J. T. P., 1989, System Indentification Methods for Faculty Detection and Diagnosis, In Proceedings of the 5th International Conference on Structural Safety and Reliability (ICOSSAR 89), San Francisco, CA. Oehler, L.T., 1957, Vibration Susceptibilities of Various Highway Bridge Types, ASCE Journal of the Structural Division, Vol. 83, pp 141. Omenzetter, P., Brownjohn, J. M. W. and Moyo, P., 2003, Identification of Unusual Events in Multi-Channel Bridge Monitoring Data Using

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

414

The Shock and Vibration Digest / September 2006 Shelley, S. J, Lee, K. L., Aksel, T. and Aktan, A. E., 1995, Active Control and Forced Vibration Studies on a Highway Bridge, ASCE Journal of Structural Engineering, Vol. 121(9), 13061312. Shepard, R. and Charleson, A. W., 1971, Experimental Determination of the Dynamic Properties of a Bridge Substructure, Bulletin of the Seismological Society of America, Vol. 61, 15291548. Shepard, R. and Sidwell, G. K., 1973, Investigations of the Dynamic Properties of Five Concrete Bridges, In Proceedings of the Fourth Australian Conference on the Mechanics of Structures and Materials, University of Queensland, Brisbane, Australia, 261268. Shepard, R., Brown, H. E. E. and Wood, J. H., 1979, Dynamic Investigations of the Mohaka River Bridge, Proceedings of the Institute of Civil Engineers, Vol. 66, 457469. Shinozuka, M. and Ghanem, R., 1995, Structural System Identification II: Experimental Verification, ASCE Journal of Engineering Mechanics, Vol. 121(2), 265273. Smyth, A. W., Masri, S. F., Chassiakos, A. G. and Caughey, T. K., 1999, Online Parametric Identification of MDOF Nonlinear Hysteretic Sytems, ASCE Journal of Engineering Mechanics, Vol. 125(2), 133142. Sohn, H., Dzwonczyk, M., Straser, E. G., Kiremidjian, A. S., Law, K. H. and Meng, T., 1999, An Experimental Study of Temperature Effect on Modal Parameters of the Alamosa Canyon Bridge, Earthquake Engineering and Structural Dynamics, Vol. 28, 879897. Srinivasan, M. G., Kot, C. A., Hsieh, B. J. and Chung, H. H., 1981, Dynamic Testing of As-Built Power Plant Buildings: An Evaluation Review, Nuclear Engineering and Design, Vol. 66, 97115. Stephen, G. A., Brownjohn, J. M. W. and Taylor, C. A., 1993, Measurements of Static and Dynamic Displacement from Visual Monitoring of the Humber Bridge, Engineering Structures, Vol. 15(3), 197208. Stubbs, N., Kim, J.-T. and Topole, K., 1992, An Efficient and Robust Algorithm For Damage Localization in Offshore Platforms, In Proceedings of ASCE Tenth Structures Congress, pp. 543546. Stubbs, N., Kim, J.-T. and Farrar, C.R., 1995, Field Verification of a Nondestructive Damage Localization and Severity Estimation Algorithm, In Proceedings of the 13th International Modal Analysis Conference (IMAC 13), Nashvile, TN, 210218. Tanaka, H. and Davenport, A. G., 1983, Wind-Induced Response Golden Gate Bridge, ASCE Journal of Engineering Mechanics, Vol. 109(11), 296312. Teughels, A. and De Roeck, G., 2004, Structural Damage Identification of the Highway Bridge Z24 by FE Model Updating, Journal of Sound and Vibration, Vol. 278(3), 589610. Thoman, S. J., Redfield, C. M. and Hollenbeck, R. E., 1984, Load Test of Single-Cell Box Bridge Cantilever Deck, ASCE Journal of Structural Engineering, Vol. 110, 17731785. Toksoy, T. and Aktan, A. E., 1994, Bridge Condition Assessment by Modal Flexibility, Experimental Mechanics, Vol. 34(3), 271278. Turer, A. and Aktan, A. E., 1999, Issues in Super-Load Crossing Of Three Steel-Stringer Bridges in Toledo, Ohio, Transportation Research Record (TRB), Issue 1688, 8796. Turer, A., Levi, A. and Aktan, A. E., 1998, Instrumentation, Proof-Testing and Monitoring of Three Reinforced Concrete Deck-on-Steel Girder Bridges Prior to, During, and After Superload, University of Cincinati, Report No. UC-CII 981. Van Nunen, J. W. G. and Persoon, A. J., 1982, Investigation of the Vibrational Behavior of a Cable-Stayed Bridge under Wind Loads, Engineering Structures, Vol. 4, 99105. Varney, R. F. and Galambos, C. F., 1966, Field Dynamic Loading Studies of Highway Bridges in the US 19481965, Transportation Research Record (TRB), no. 76, 285304. Ventura, C. E., Felber, A. J. and Stiemer, S. F., 1995, Experimental Investigations of Dynamics of Queensborough Bridge, Journal of Performance of Constructed Facilities, Vol. 9(2), 146155. Ventura, C. E., Felber, A. J. and Stiemer, D. F., 1996 Determination of the Dynamic Characteristics of the Colquitz River Bridge by Full-Scale Testing, Canadian Journal of Civil Engineering, Vol. 23(2), 536548. Vincent, H. T., Hu, S. J. and Hou, Z., 1999, Damage Detection Using Empirical Mode Decomposition Method and a Comparison with Wavelet Analysis, In Proceedings of the 2nd International Workshop on Structural Health Monitoring, Stanford University, Palo Alto, CA, 891900. Wahab, M. M. A. and De Roeck, G., 1998, Dynamic Testing of Prestressed Concrete Bridges and Numerical Verification, Journal of Bridge Engineering, Vol. 3(4), 159169. Ward, H. S., 1984, Traffic Generated Vibrations and Bridge Integrity, ASCE Journal of Structural Engineering, Vol. 110, 24872498.

Wavelet Transform and Outlier Analysis, Proceedings of SPIE The International Society for Optical Engineering, Vol. 5057, 157 168. Omenzetter, P., Brownjohn, J. M. W. and Moyo, P., 2004, Identification of Unusual Events in Multi-Channel Bridge Monitoring Data, Mechanical Systems and Signal Processing, Vol. 18(2), 409430. Pardoen, G. C., Carr, A. J. and Moss, P. J., 1981, Bridge Modal Identification Problems, In Proceedings of the Second Specialty Conference on Dynamic Response of Structures: Experimental, Observation, Prediction and Control, 2945. Peeters, B. and De Roeck, G., 2000, Reference Based Stochastic Subspace Identification In Civil Engineering, Inverse Problems in Engineering, Vol. 8(1), 4774. Peeters, B. and De Roeck, G., 2001, One-Year Monitoring of the Z24Bridge: Environmental Effects Versus Damage Events, Earthquake Engineering and Structural Dynamics, Vol. 30, 149171. Peeters, B. and Ventura, C. E., 2003, Comparative Study of Modal Analysis Techniques for Bridge Dynamic Characteristics, Mechanical Systems and Signal Processing, Vol. 17(5), 965988. Peeters, B., Maeck, J. and De Roeck, G., 2001, Vibration-Based Damage Detection in Civil Engineering: Excitation Sources and Temperature Effects, Smart Materials and Structures, Vol. 10, 518527. Penrose, R., 2005, The Road to Reality, A Complete Guide to the Laws of the Universe, Alfred A. Knopf, New York, NY. Piombo, B. A. D., Fasana, A., Marchesiello, S. and Ruzzene, M., 2000, Modeling and Identification of the Dynamic Response of a Supported Bridge, Mechanical Systems and Signal Processing, Vol. 14(1), 7589. Pridham, B. A. and Wilson, J. C., 2002, Subspace Identification of Vincent Thomas Suspension Bridge Ambient Vibration Data, In Proceedings of the 20th International Modal Analysis Conference (IMAC 20), Los Angeles, CA, 134140. Pridham, B. A. and Wilson, J. C. 2005, A Reassessment of Dynamic Characteristics of the Quincy Bayview Bridge Using Output-Only Identification Techniques, Earthquake Engineering and Structural Dynamics, Vol. 34(7), 787805. Radkowski, A. F., Bakht, B. and Billing, J., 1984, Design and Testing of a 400 ft. Span Plate Girder Bridge, In Proceedings of US/Japan Joint Seminar on Composite and Mixed Construction, Seattle, WA, 6070. Raghavendrachar M. and Aktan, A. E., 1992, Flexibility by Multireference Impact Testing for Bridge Diagnostics, ASCE Journal of Structural Engineering, Vol. 118(8), 21862203. Ren, W. X., Zhao, T. and Harik, I. E., 2004, Experimental and Analytical Modal Analysis of Steel Arch Bridge, ASCE Journal of Structural Engineering, Vol. 130(7), 10221031. Richardson, J. A. and Douglas, B. M., 1987, Identifying Frequencies and Three-Dimensional Mode Shapes from a Full Scale Bridge Test, In Proceedings of the 5th International Modal Analysis Conference on Structural Dynamics (IMAC 5), London, UK, 160165. Richardson, J. A., and Douglas, B. M., 1993, Results from Field Testing a Curved Box Girder Bridge Using Simulated Earthquake Loads, Earthquake Engineering and Structural Dynamics, Vol. 22(10), 905922. Roberts, G. P. and Pearson, A. J., 1998, Health Monitoring of StructuresTowards a Stethoscope for Bridges, In Proceedings of the International Conference on Noise and Vibration Engineering (ISMA 23), Leuven, Belgium. Rohrmann, R. G., Baessler, M., Said, S., Schmid, W. and Rucker, W. F., 2000, Structural Causes of Temperature Affected Modal Data of Civil Structures Obtained by Long Time Monitoring, In Proceedings of the 18th International Modal Analysis Conference on Structural Dynamics (IMAC 18), San Antonio, TX, 17. Saiidi, M. and Douglas B. M., 1984, Effect of Design Seismic Loads on Highway Bridge, ASCE Journal of Structural Engineering, Vol. 110, 27232737. Salane, H. J., Baldwin, J. W. and Duffield, R. C., 1981, Dynamics Approach for Monitoring Bridge Deterioration, Transportation Research Record (TRB), Issue 832, 2128. Sartor, R., Culmo, M.P. and DeWolf, J. T., 1999, Short Term Strain Monitoring of Bridge Structures, ASCE Journal of Bridge Engineering, Vol. 4(3), 157164. Scalon, A. and Mikhailovsky, L., 1987, Full-Scale Load Test of ThreeSpan Concrete Bridge, Canadian Journal of Engineering, Vol. 14, 1923. Shahrooz, B. M., Ho, I. K., Aktan, A. E., Borst, R., Blaauwendraad, J., Veen, C., Iding, R. H. and Miller, R. A., 1994, Nonlinear Finite Element Analysis of a Deteriorated Slab Bridge, ASCE Journal of Structural Engineering, Vol. 120(2), 422440.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS Wenzel, H. and Pichler, D., 2005, Ambient Vibration Monitoring, John Wiley & Sons, NJ. White, J. M. and Pardoen, G. C., 1987, Modal Identification of the Golden Gate Bridge Tower Using Ambient Vibration Data, In Proceedings of the 5th International Modal Analysis Conference on Structural Dynamics (IMAC 5), London, UK, 1620. Wilson, J. C. and Lui, T., 1991, Ambient Vibration Measurements on a Cabled-Stayed Bridge, International Journal of Earthquake Engineering and Structural Dynamics, Vol. 20(8), 723748. Wong, H. L., Trifunac, M. D. and Luco, J. E., 1988, Comparison of SoilStructure Interaction Calculations with Results of Full-Scale Forced Vibration Tests, Soil Dynamics and Earthquake Engineering, Vol. 7(1). Wong, K. Y., Chan, W. Y. K., Man, K. L., Mak, W. P. N. and Lau, C. K., 2000, Structural Health Monitoring Results on Tsing Ma, Kap Shui Mun and Ting Kau Bridges, In Proceedings of SPIE Nondestructive Evaluation of Highways, Utilities and Pipelines IV, Newport Beach, CA, Vol. 3995, 288299. Wong, K. Y., Lau, C. K. and Flint, A. R., 2000, Planning and Implementation of the Structural Health Monitoring System for Cable-Supported Bridges in Hong Kong, In Proceedings Of SPIE Nondestructive Evaluation of Highways, Utilities and Pipelines IV, Newport Beach, CA, Vol. 3995, 266275. Wong, K. Y., Chan, K. W. Y. and Man, K. L., 2001, Monitoring of Wind Load and Response for Cable-Supported Bridges in Hong Kong, In Proceedings Of SPIE Health Monitoring and Management of Civil Infrastructure Systems, Newport Beach, CA, Vol. 4337, 292303. Xia, P.-Q. and Brownjohn, J. M. W., 2003, Residual Stiffness Assessment of Structurally Failed Reinforced Concrete Structure by Dynamic Testing and Finite Element Model Updating, Experimental Mechanics, Vol. 43(4), 372378. Yang, J. N., Lei, Y. and Huang, N. E., 2001, Damage Identification of Civil Engineering Structures Using Hilbert-Huang Transform In Proceedings of the 3rd International Workshop on Structural Health Monitoring, Stanford University, Palo Alto, CA, 544553. Yi, T., Moon, F. L., Leon, R. T., and Kahn, L. F., 2006 Lateral Load Tests on a Two-Story Unreinforced Masonry Building, ASCE Journal of Structural Engineering, Vol. 132(5). Yuen, K.-V. and Katafygiotis, L. S., 2002, Bayesian Modal Updating Using Complete Input and Incomplete Response Noisy Measurements, Journal of Engineering Mechanics, Vol. 128(3), 340350. Zapico, J. L., Gonzalez, M. P. and Worden, K., 2003, Damage Assessment Using Neural Networks, Mechanical Systems and Signal Processing, Vol. 17(1), 119125. Zhang, R. and Aktan, A. E., 2005, Design Considerations for Sensing Systems to Ensure Data Quality, In F. Ansari (Ed.) Sensing Issues in Civil Structural Health Monitoring, Springer, Dordrecht, The Netherlands, 281291. Zhang, R., Aktan, A. E, Oh, P., Reddy, A. and Wen, J., 2004, Unified Sensor Network Design for Intelligent Infrastructure Systems, Proposal 0427946 submitted to NSF 04522. Zhao, J. and DeWolf, J. T., 2002, Dynamic Monitoring of Steel Girder Highway Bridge, ASCE Journal of Bridge Engineering, Vol. 7(6).

415

were made to stand underneath their completed structures as they were loaded according to design requirements. The importance of field testing continued to be recognized up to modern times, and according to Billington (1983) some of the most prolific structural engineers practicing around the turn of the 20th Century relied on full-scale tests and observations of system behavior to develop, validate and augment theory. For instance, in 1907 Robert Maillart argued that designers should be encouraged to check their assumptions through load tests that establish deflections of the completed structure (Billington, 1979). The contributions of such pioneering efforts notwithstanding, it is important to draw a distinction between this type of testing and the rigorous scientific experimentation, enabled by advances in technology, that first began to appear in an elementary form as early as the 1950s. In order to categorize the large amount of studies identified, the organizational structure shown in Table 4 is used. The broadest distinction included in this approach is related to the type of excitation employed, and classifies studies as either controlled or uncontrolled. These terms refer simply to whether the excitation was intentionally provided, and do not reflect whether it was quantified or not (e.g., a crawl speed truck load test would be considered controlled excitation). Additional distinctions are related to the level of testing (i.e., non-destructive or destructive) for controlled tests and the source of excitation (i.e., ambient [wind, traffic, etc.] or earthquake) for uncontrolled tests. Finally, each test is categorized based on the structure type investigated. The following sections aim to provide a very broad brush treatment of the past experimental studies on actual (operating or decommissioned) constructed facilities that fall into the categories highlighted in Table 4. These categories were chosen to provide the reader with a sufficient background to insure that the discussions presented in previous sections, which touch on lessons learned and future challenges, are taken in context. While substantial contributions to field testing have been made by studies that reside in the other categories, a review of this work is left to others. A.1. Controlled, Nondestructive Experiments A.1.1 Short- to Moderate-Span Bridges The first pioneering work carried out in the mold of todays scientific research was focused on bridge structures, and can be characterized as nondestructive, controlled testing. Varney and Galambos (1966) and Oehler (1957) report numerous dynamic tests of bridges that employed vehicles for dynamic excitation and were conducted throughout the

Appendix: Chronology of Field Experimentation To the authors knowledge, the first reported practice of testing constructed facilities within their operating environments can be traced back to the ancient Romans in the first Century B.C. Intending to ensure diligent practice, architects

Table 4. Proposed organization chart for field studies of constructed facilities. Loading Type Controlled Classification Nondestructive Destructive Uncontrolled Ambient Earthquake Short- to Moderate-Span Bridges Section A.1.1 Section A.2 Section A.3.1 X Structure Type Long-Span Bridges Section A.1.2 X Section A.3.2 X Buildings and Other Structures X X X X

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

416

The Shock and Vibration Digest / September 2006

US between 1948 and 1965. In addition, Iwasaki et al. (1972) reviewed 26 dynamic tests of highway bridges in Japan (using eccentric mass shakers and rocket engines for excitation) carried out from 19581969, and Cantieni (1984) summarized 226 dynamic tests of bridges (using test vehicles, shakers, and impact for excitation) conducted in Switzerland between 1958 and 1981. A principal outcome of these last two investigations was an empirical relationship for approximating resonant frequencies and damping ratios based on span length. Using both test vehicles and eccentric mass shakers, Shepard and Sidwell (1973) tested five reinforced concrete bridges and noted that the dynamic properties compared well with analytically predicted values. In addition, the authors reported that the horizontal modes have considerably higher damping than the vertical modes, which they attributed to soilstructure interaction. Additional studies that utilized eccentric mass shakers throughout the 1970s include: Shepard and Charleson (1971), who reported the vibration testing of a continuous bridge at various stages of construction; Kuribayashi and Iwasaki (1973), who detailed the forced vibration testing of 30 highway bridges; Leonard and Eyre (1975), who discussed testing procedures and equipment; and Shepard et al. (1979), who reported the results of forced vibration testing of a three-span steel truss bridge. The step-relaxation excitation approach was utilized by Douglas (1976) and Douglas et al. (1981) to identify resonant frequencies and damping ratios for a composite bridge and a reinforced concrete bridge, respectively. To investigate whether changes in dynamic properties could be used to assess damage, Salane et al. (1981) conducted both transient and steady-state dynamic testing using an electro-hydraulic actuator system during the fatigue testing of a three-span highway bridge. Although the experiments discussed thus far provided important information about the state of their respective test bridges, it is important to point out that these studies simply aimed to identify the dynamic properties of the structure. In contrast, Douglas and Reid (1982) were the first to employ a structural identification procedure to properly identify an analytical model based on vibration measurements of a bridge. Using the step-relaxation excitation method, the authors identified transverse dynamic properties and used these to identify a model that provided a better understanding of the dynamic characteristics of the bridge, as well as information about soil-structure interaction. In another study, Buckle et al. (1986) and Richardson and Douglas (1993) report a dynamic test of a curved concrete box girder bridge located in New Zealand. The bridge was excited using the step-relaxation technique and a dynamic response model was calibrated based on the identified fundamental frequency and mode shape through the use of a structural identification procedure. This model was able to provide information regarding the stiffness of the foundation, bearings, and concrete. In addition, Natke and Yao (1986, 1989) provide an overview of various considerations associated with structural identification and discuss critical research areas, including damage detection and the identification of nonlinearities among others. While these studies illustrated the many advantages of employing structural identification procedures for behavior studies, several investigators continued to focus on more narrow topics such as the identification of dynamic properties (Billing, 1984; Richardson and Douglas, 1987) the validation of ana-

lytical or design models (Thoman et al., 1984; Radkowski et al., 1984; Crouse et al., 1987; Lee et al., 1987; Gilani et al., 1995) and the response of bridges to seismic design loads (Saiidi and Douglas, 1984). The 1990s saw a dramatic increase in the number of studies that employed structural identification procedures. Raghavendrachar and Aktan (1992) employed structural identification in conjunction with multiple reference impact testing to investigate integrity monitoring and diagnostics of bridges. The authors present a method for transforming mode shape coefficients directly into flexibility information without having to make any assumption of mass. The identified analytical model was used to demonstrate that flexibility coefficients are more sensitive to local damage than either frequencies or mode shapes. Throughout the mid-1990s a series of studies were carried out to investigate the use of structural identification to develop an improved understanding of actual load resisting mechanisms of common bridge types (Aktan et al., 1993a; Aktan et al., 1995a, 1995b; Aktan et al., 1996; Farhey et al., 1997; Lenett et al., 1997; Catbas et al., 1997; Turer et al., 1998). These studies employed multiple reference impact tests to develop modal flexibility models, which were subsequently validated through static crawlspeed truck tests. Once validated, these response measurements were used to carry out comprehensive structural identification of 3-D finite element models. The authors noted that by incorporating the identified load carrying mechanisms (e.g., composite action for bridges designed to act non-compositely) into the analytical model, excellent correlation with the experiment was achieved. Many benefits of integrating this type of rigorous experimentation with analytical simulation models were noted. First, bridge-rating factors obtained using the identified models were several times (up to 4.0) larger than those obtained by idealized models (Aktan et al., 1995a). Second, these models permitted the best design for superload axle configuration, crossing paths and positions on bridges, as well as possible options for effective bridge strengthening when needed (Turer et al., 1999). In addition, Shelley et al. (1995) employed identified models to aid in developing an active vibration control approach that was demonstrated to reduce the global mode response of a truss bridge by 90%. Lastly, it was noted that experimental information contributed to an increase in the level of confidence in the condition assessment of a bridge and resulted in a more economical retrofit design (Farhey et al., 2000). Taken as whole, these studies demonstrated that employing structural identification with field experimentation greatly enhances both the reliability and the benefit-to-cost ratio of the research. In an attempt to generalize these research results, Aktan et al. (1997a, 1998a) provided overviews of the experimental arts and analytical aspects of structural identification, respectively. Farrar et al. (2000) report the dynamic testing of the Alamosa Canyon Bridge a seven-span composite bridge) using multiple reference impact and shaker excitation as well as ambient excitation. A primary focus of this effort was to investigate the variability of modal properties due to environmental effects, vehicle weight, excitation source, and data reduction process. Bootstrap analyses revealed that variability associated with the measurement and data reduction processes were not as significant as the variability produced by the temperature difference across the deck, which was

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS

417

found to cause variations in modal frequencies of up to 6% over a 24-hour period (Farrar et al., 1997; Cornwell et al., 1999). To investigate the impact of this variability on damage detection, Doebling et al. (1997) developed 95% confidence intervals of the modal properties based on measured data and used an identified finite element model to investigate the impact of several damage scenarios on modal parameters. The authors concluded that modes to be used for damage detection should be carefully selected based on the inherent uncertainty caused by environmental effects as well as their sensitivity to structural deterioration. Sohn et al. (1999) developed a linear adaptive model to distinguish changes in modal parameters caused by temperature gradients from those caused by other effects. This model was then used to establish confidence intervals of modal parameters based on new temperature profiles. As part of an island-wide bridge upgrading in Singapore, Moyo et al. (2004) conducted forced vibration testing (shaker excitation) and ambient vibration monitoring to assess the effect of upgrading measures on an 18 meter span bridge. The primary upgrading procedure involved replacing nominally pinned supports with nominally fixed supports, which resulted in a very large increase in bridge stiffness (Brownjohn et al., 2002). By employed structural identification procedures to develop calibrated finite element models from data obtained before and after retrofit, Brownjohn et al. (2003) showed that the alteration of boundary conditions was the principal cause of increases (of up to 50%) in modal frequency. The authors note that combining the experimental and analytical processes by direct identification of structural properties is required to obtain such results (and confidence) and point out that this procedure, in general, can be applied to other structures. The studies discussed above represent perhaps the most influential controlled, nondestructive investigations since the early 1990s that have specifically demonstrated the benefits of employing structural identification in conjunction with field experimentation. However, several other experimental studies of constructed facilities have been carried out and merit mentioning. Green and Cebon (1993) discussed the results of a multiple reference impact test of a four-span continuous prestressed concrete box girder bridge. Using a servo-hydraulic shaker, Cantieni and Pietrzko (1993) conducted forced vibration testing of a timber footbridge to investigate potential vibration problems. Kohoutek (1993) reported the results of static and dynamic testing of the Talbragar River Bridge. Through testing of the Colquitz River Bridge, Ventura et al. (1996) showed the consistency of modal parameters identified through step-relaxation excitation and ambient testing. Wahab and De Roeck (1998) investigated the correlation between finite element models and experimental vibration testing using drop weight, test vehicle, and ambient excitation of three bridges in Belgium. Christensen et al. (2001) and Halling et al. (2004) conducted forced and ambient excitation tests of bridges along the I-15 corridor in Salt Lake City, Utah. These studies investigated the influence of forcing amplitude/ direction and temperature on the modal parameters. Using a structural identification procedure, Xia and Brownjohn (2003) were able to determine the residual stiffness distribution along a failed reinforced concrete beam structure. Finally, Culmo et al. (2004) compared strain gage data taken during the crossing of a 1,000,000 lb permit vehicle with estimates of live load stresses produced using structural analysis.

A.1.2 Long-Span Bridges Compared to mid- and short-span bridges, controlled testing is far less frequently utilized for testing long span bridges due to their higher flexibility, which presents two general challenges. First, the ambient vibrations of these structures tend to have relatively large amplitudes and thus it is difficult to overcome these vibrations with controlled excitation. Second, the modal frequencies of long-span bridges can be quite low (~1 Hz) and as a result it is difficult to supply forced excitation (e.g., through the use of a shaker) that has sufficient energy at such low frequencies. However, despite these challenges, several researchers have employed controlled testing for long-span bridges. Buckland et al. (1979) reported the vibration testing of a suspension bridge using a weighted pendulum, vehicle braking, and a vehicle driving over timber boards for excitation. Experimental results compared well with analytical predictions (except for torsional modes) and the authors noted strong coupling between vertical and horizontal modes. Using both step-relaxation and ambient excitation Muria-Vila et al. (1991) conducted vibration testing on a three-span cable-stayed bridge. Results indicated good agreement between controlled and ambient excitation tests. Filiatrault et al. (1993a, 1993b) tested the earthquake-damaged Shipshaw cable-stayed bridge using test vehicles at constant speed and through vehicle braking. The authors point out difficulties identifying damping ratios, but report good correlation with a finite element model. Finally, Fujino et al. (1999, 2000a) report the forced and ambient vibration testing of the Hakucho suspension bridge. Forced excitation was supplied by two vibrating machines placed at a quarter-point and mid-point of the main span. Results indicated that the fundamental frequency and mode shape (first bending mode) were dependant on wind speed. The authors conclude that this dependency may be related to friction in the bearings at the towers. A.2. Controlled, Destructive Experiments While the contributions of nondestructive field tests are undeniable, as early as the 1970s researchers began to explicitly identify the need for destructive testing (Hudson, 1977). The following sections present a brief discussion of this type of testing for short- to moderate-span bridges, as no destructive tests of long-span bridges were identified. Jorgenson and Larson (1972) tested a reinforced concrete bridge to collapse using hydraulic rams. The authors noted that the measured response compared well to design calculations at the proportional limit and ultimate loads. In addition, Burdette and Goodpasture (1972) tested four highway bridges to failure and noted that the AASHTO specifications provided lower bounds of bridge ultimate capacity. McClure and West (1984) and Scalon and Mikhailovsky (1987) report the results of full-scale tests of a prestressed concrete bridge and a three-span continuous reinforced concrete bridge, respectively. Following the Loma Prieta earthquake, Bollo et al. (1990) conducted forced vibration and destructive tests, before and after retrofit, of the I-880 Cypress Viaduct in Oakland, CA. These tests quantitatively demonstrated the effectiveness of the retrofits and illustrated that accurate predictions of frequencies, mode shapes, and damping depended to a large degree on proper modeling of the foundation. The

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

418

The Shock and Vibration Digest / September 2006

nondestructive and destructive testing of a decommissioned, skewed reinforced concrete flat slab bridge and associated analytical studies was reported Miller et al. (1992, 1994), Aktan et al. (1993b), Toksoy and Aktan (1994), and Sharooz et al. (1994). The goal of this study was to investigate the use of nondestructive evaluation procedures to assess bridge condition. Multiple reference impact testing was carried out at various levels of damage induced through the application of static vertical loads. Damage was located and quantified through the use of both modal flexibility coefficients and FE models calibrated through the use of a structural identification procedure. Results indicated the excessive conservatism of current bridge rating models, as the ultimate capacity of the bridge was approximately four times greater than predicted by such models (equivalent to 22 rating trucks). To investigate the performance of historic truss bridges, Aktan et al. (1993b, 1995b) tested two 80-year old through truss bridges to failure using hydraulic jacks supported by rock anchors. During intermediate stages each bridge was evaluated nondestructively using an eccentric mass shaker, linear shaker, and crawl-speed truck tests. To obviate local failures, the trusses were retrofitted by welding steel plates to strengthen certain critical connections. The authors concluded that both bridges displayed adequate performance at the various limit states investigated and possessed sufficient deformability to allow for extensive force redistribution. Lauzon and DeWolf (1995) monitored bridge vibrations during incrementally induced cuts of the bottom flange and web of a decommissioned bridge and concluded that major defects cause changes in a bridges dynamic signature. Eberhard and Marsh (1997a,b) imposed quasi-static loads on a three-span continuous reinforced concrete bridge and isolated reinforced concrete bridge bents, respectively. Results indicated that approximately 90% of the lateral bridge stiffness was due to longitudinal confinement supplied by the abutments. In addition, despite the 1960-era, pre-seismic design, details, the bridge bents exhibited ductile behavior, which was attributed to the rotational flexibility of the soil. To assess shear capacity of reinforced concrete T-beam bridges, a decommissioned bridge was loaded to failure by Al-Mahaidi et al. (2000). Finally, Halling et al. (2001) performed several forced vibration tests on a single-span of a freeway overpass in conjunction with a series of intentionally induced damage states and repairs. Through the use of a structural identification procedure, the authors calibrated analytical models and demonstrated that such models are capable of successfully identifying and estimating the location and extent of damage. Aside from the studies outlined thus far, a small group of destructive tests deserve special attention due to their extensive documentation, and, in some cases, the participation of a large portion of the health monitoring community. The first such investigations involved the destructive and nondestructive testing of the I-40 Bridge over the Rio Grande and associated numerical studies (Farrar et al., 1994, 1996; Mayes and Nusser, 1994). The investigation began with a series of ambient vibration tests using traffic (both on the I-40 Bridge and on an adjacent bridge) for excitation and forced vibration tests using a vertical linear mass shaker. The authors report that the results from these tests compared well. However, it was noted that the various ambient vibration tests showed a 510% variation in frequencies and mode shapes, which was attributed to temperature fluctuations. Following

this initial testing, the bridge was intentionally damaged by introducing four levels of cuts in one of the two plate girders. Following each level of damage the bridge was subjected to additional vibration tests to determine any changes in modal properties. Results indicated that only during the largest damage case (when one of the two plate girders was cut halfway through) was there an appreciable change in natural frequencies, and this change was comparable to that observed during the various ambient tests and attributed to temperature changes. In addition to demonstrating the difficulty of assessing damage in a real civil structure, the data from the I-40 Bridge was exploited by several investigators to investigate the efficacy of the various damage detection algorithms (Mayes, 1995; Stubbs et al., 1995; Farrar and Jauregui, 1996, 1998a,b) and statistical approaches (Farrar et al., 2000). To investigate vibration-based damage detection and to compare objective and subjective methods of bridge condition evaluation, a steel stringer/reinforced concrete bridge (the Seymour Avenue Bridge) in Cincinnati, OH was subjected to a series of nondestructive tests in which several damage scenarios were induced (Aktan et al., 1997b; Catbas et al., 1998, 2004; Lenett et al., 1999; Catbas and Aktan, 2002). The damage scenarios included jammed bearings, removal of a bearing, cuts to various supporting members, and delaminating a portion of the slab from the girders. Following each damage scenario, the bridge was subjected to multi-reference impact testing and crawl-speed truck tests. The results of these tests were compared to evaluate the consistency of modal flexibility and static flexibility. Results indicated that modal flexibility is an accurate measure of in-situ flexibility and a reliable index for bridge condition and damage assessment. Although the bridge was not loaded to failure, the results indicated that bridges capacity was much greater than indicated by current bridge rating models. The authors concluded that subjective methods of bridge evaluation and assessment are not able to characterize intrinsic bridge mechanisms and their influence on bridge behavior. To study the effects on the dynamic properties of environmental effects and damage, a three-span post-tensioned box girder bridge (Z24 Bridge) in Canton Bern, Switzerland was tested by EMPA (Swiss Federal Laboratories for Materials Testing and Research) (DeRoeck et al., 2000; DeRoeck, 2003). The testing consisted of a one-year monitoring study to quantify the effect of environmental influences and series of nondestructive tests carried out as the structure was progressively damaged. Damage scenarios included foundation settlement, a landslide, failure of concrete hinges, and rupture of post-tensioning tendons, among others. Following the introduction of each damage scenario, the bridge was subjected to both ambient and forced vibration testing to determine its dynamic properties and assess if the observed changes were statistically significant. Based on the results, the authors concluded that reliable modal information can be obtained from output only tests, and that damage has a selective influence on modes. Numerous damage detection studies have been carried out using the data obtained from this experiment both by researchers involved in the project (e.g., Abdel Wahab and DeRoeck, 1999; Maeck et al., 2001; Peeters et al., 2001; Maeck and DeRoeck, 2003; Teughels and De Roeck, 2004), and by many others. In addition, this data was made available as a benchmark study aimed at

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

F.L. Moon and A.E. Aktan / STRUCTURAL IDENTIFICATION OF CONSTRUCTED (CIVIL) SYSTEMS

419

comparing modal analysis techniques for evaluating the dynamic response data obtained from forced, free and ambient vibration tests. These studies were included as a special session at the International Modal Analysis Conference in 2001 (Peeters and Ventura, 2003). A.3. Uncontrolled, Ambient Experiments The following section briefly describes past tests that have exclusively employed ambient excitation to perform vibration tests of bridges. The numerous investigations that employed ambient vibration tests in conjunction with controlled testing have been outlined in the previous sections. A.3.1 Short- to Moderate-Span Bridges Early ambient vibration studies include Pardoen et al. (1981), who reported the results an ambient vibration test of a steel truss bridge, Gates and Smith (1982), who summarize the results of ambient vibration tests on 57 highway bridges in California, and Ward (1984), who reports on the ambient vibration tests of 18 highway bridges, primarily using traffic for excitation. In addition to case specific results, these studies identified two fundamental challenges related to ambient testing: (1) the nature of input may cause peaks in the Fourier spectra that do not represent a structures modes, and (2) if long windows of data are taken, the structure itself will undergo changes due to its non-stationary nature. More recently, DeWolf et al. (1997), Bernard et al. (1997), and Sartor et al. (1999) demonstrated the usefulness of ambient monitoring studies as an aid to the fatigue evaluation of steel bridges. These authors noted that the principal motivation for such studies was the presence of stress concentrations and structural redundancy which precluded the accurate prediction of localized stresses by traditional analysis techniques. Asmussen et al. (1998a,b) performed ambient vibration testing, using the random decrement technique, of a two-span reinforced concrete bridge, to investigate the potential of vibration-based bridge inspection. To investigate the use of the wavelet estimation technique, Piombo et al. (2000) subjected a simply-supported bridge to ambient vibration tests using traffic as excitation. This study represented the first attempt to employ the wavelet estimation technique directly on transient data, rather than in conjunction with the random decrement technique. Fu and DeWolf (2001) reported the results of an ambient vibration study conducted over a period of 1 year on a slightly skewed, two-span continuous plate girder bridge with a concrete deck. This monitoring tracked the first three frequencies, which experienced increase of 1013% as the temperature dropped. This data was then used by Zhao and DeWolf (2002) to show that modified modal flexibility was more sensitive to changes in bridge response than frequencies or mode shapes. Bolton et al. (2002) report the ambient monitoring of a two-span reinforced concrete highway bridge and the automation of modal property extraction. Hales and Halling (2002) report the results of the first four months of a long-term ambient vibration monitoring study, using permanent strong-motion instrumentation, aimed at determining an accurate statistical baseline of the bridge dynamics. Omenzetter et al. (2003, 2004) employed a wavelet-based outlier detection method to identify abrupt and anomalous events in the response of

the continually monitored Tuas Link that connects Singapore and Malaysia. The proposed method was successfully validated through the use of construction events and was also successful in identifying anomalous post-construction events. Ren et al. (2004) discuss the ambient vibration testing, using traffic and wind excitation, of a steel-girder arch bridge. The authors noted good agreement between frequencies identified by peak picking and the stochastic subspace identification (SSI) method; however, the SSI method provided superior mode shapes. Finally, Kim et al. (2005) present a new technique, termed the time domain decomposition (TDD) method, to extract modal parameters from out-put only data. The feasibility of this approach was demonstrated using data from the King Storm Water Bridge located in Southern California. A.3.2 Long-Span Bridges Because of the fundamental challenge of exciting longspan bridges through forced excitation, ambient vibration methods have been widely employed on such bridges. The first study that used ambient vibration data to characterize the dynamic properties of a bridge was reported by McLamore et al. (1971). This investigation focused on two suspension bridges, and extracted frequencies and mode shapes using power and cross spectra, estimating the damping using the half-power bandwidth method. Abdel-Ghaffar and Housner (1978) investigated the extraction of damping values using ambient vibration tests on the Vincent-Thomas Suspension Bridge. Due to closely spaced spectral peaks and spectral overlap, the authors concluded that the half-power bandwidth method does not provide reliable damping values. During the design phase of a cable-stayed bridge, Van Nunen and Persoon (1982) carried out wind tunnel studies of a two-dimensional model and later compared the results with measured wind-induced vibration of the completed structure. Tanaka and Davenport (1983) and Abdel-Ghaffar and Scanlon (1985a) studied the ambient response of the suspended structure of the Golden Gate Bridge. Using the same approach as McLamore et al. (1971), these studies identified the bridges dynamic properties and noted a good agreement with analytical models. Additional studies by Abdel-Ghaffar and Scanlon (1985b) and White and Pardoen (1987) reported ambient vibration testing of a tower of the Golden Gate Bridge. Brownjohn et al. (1986, 1987) conducted an ambient vibration study of the Humber Suspension Bridge to determine the vertical, lateral and torsional vibration characteristics. This study was followed up by an extensive monitoring investigation from 19901991, to validate analytical models of the bridge and investigate the relationship between input and response parameters (Stephen et al., 1993; Brownjohn et al., 1994). The first and second Bosporus Suspension Bridges were tested under ambient vibrations by Brownjohn et al. (1989, 1992a, 1992b). A detailed comparison of experimental results and analytical models was carried out and showed reasonable agreement except for lateral modes, which was attributed to the low degree of excitation. In addition, the authors point out the inherent error associated with assuming that ambient input is random and has a flat power spectrum. To investigate the wind response of flexible bridges, Kumarasena et al. (1989, 1991) conducted ambi-

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

420

The Shock and Vibration Digest / September 2006

ent vibration monitoring of the Deer Isle Bridge in Maine. These studies illustrated the effectiveness of cable-bracing in reducing wind-induced vibrations and helped confirm that self-induced turbulence is a primary mechanism of the buffeting-type response. Wilson and Lui (1991) reported ambient vibration tests of a cable-stayed bridge and noted difficulty in determining damping ratios because of the nonstationary nature of the input. Ventura et al. (1995) investigated the dynamic properties of the Queensborough Bridge through ambient monitoring as part of a seismic retrofit program. The authors noted that the results, particularly related to the effects of soil-structure interaction, proved useful for the subsequent retrofit efforts. In the late 1990s, structural health monitoring systems were installed on the Tsing Ma Suspension Bridge and the Kap Shui Mun and Ting Kau cable-stayed bridges in Hong Kong, and have been operating ever since (Kwong et al., 1995; Wong et al., 2000a, 2000b). Various researchers have used this data to investigate various bridge responses (Wong et al., 2000a, 2000b, 2001; Chang et al., 2001), the feasibility of vibrationbased damage detection (Ko et al., 2000), and the empirical mode decomposition method in conjunction with the Hilberttransform technique for modal parameter identification (Chen et al., 2004). Brownjohn et al. (1999) studied the dynamic response of a 100 m span cable-stayed bridge located in Singapore. Through comparisons with analytical models, it was concluded that proper modeling of the deck end fixity was necessary to properly simulate bridge response, while capturing the stress-stiffening effect of the stay cables was of lesser importance. Abe et al. (2000) conducted an ambient vibration study on the Hakucho Suspension Bridge in northern Japan to investigate dynamic behavior and develop a

health monitoring system. Using the RD method in conjunction with the ITD method, the authors estimated modal parameters, located fatigue prone areas, and showed that the lower frequencies were highly dependent on wind speed (Mahmoud et al., 2001). Fujino et al. (2000b) discussed the use of GPS to monitor the Akashi Kaikyo Bridge in Japan. The authors indicate the potential of using GPS measurements to ascertain deformations caused by extreme events and those that fall outside the normal operating range. Brincker et al. (2000, 2001) conducted vibration tests on the Great Belt Bridge to investigate the extraction of reliable damping values from ambient response data. In this study the authors employed and compared three different methods, including the Frequency Domain Decomposition (FDD) and both data-driven and covariance-driven SSI techniques. Banish et al. (2000) and Aktan et al. (2000a) discuss the design and implementation of an extensive health monitoring system on the Commodore Barry Bridge outside of Philadelphia, PA. Using data obtained from this system, structural identification studies were carried out (Catbas et al., 2000; Aktan et al., 2000b) and the limitations of such efforts were identified (Catbas et al., 2003). Pridham and Wilson (2002) employed the SSI technique to identify modal parameters of the Vincent Thomas Suspension Bridge from ambient vibration data. Using the same bridge, Masri et al. (2004) discussed the architecture and optimization of a health monitoring system for deployment to distributed civil structures. Finally, Pridham and Wilson (2005) employed the SSI technique to extract modal parameters of the Quincy Bayview CableStayed Bridge from response data recorded in 1987. Results indicated a dramatic improvement in the estimation of damping values over earlier studies that employed spectral analysis.

Downloaded from http://svd.sagepub.com at PENNSYLVANIA STATE UNIV on April 14, 2008 2006 SAGE Publications. All rights reserved. Not for commercial use or unauthorized distribution.

Vous aimerez peut-être aussi