2012-12-26-symptom_based reliability and generalizedre pairing cost

Upload: sholran

Post on 03-Apr-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    1/9

    Symptom-based reliability and generalized repairing cost

    in monitored bridges

    R. Ceravolo , M. Pescatore, A. De Stefano

    Politecnico di Torino, Torino, Italy

    a r t i c l e i n f o

    Article history:

    Received 24 May 2007

    Received in revised form

    22 January 2009

    Accepted 11 February 2009Available online 20 February 2009

    Keywords:

    Structural reliability

    Safety assessment

    Modal testing

    Health monitoring

    Symptom

    Generalized maintenance cost

    a b s t r a c t

    This paper proposes the use of structural safety formulations conceived to take into account the

    presence of periodic monitoring systems. Monitoring is a valid tool to improve the safety of those

    structural systems that cannot withstand invasive tests or interventions that would alter their nature or

    their intended use. Reliability can be defined as a function of a measurable quantity that reflects the

    damage, referred to as symptom, and it can also be defined as a function of several symptoms

    considered simultaneously. A knowledge of the current value of a symptom makes it possible to

    determine the residual damage capacity and the residual lifetime of a structure. Redefining structural

    safety in terms of residual lifetime provides the theoretical framework for the introduction of vibration-

    based monitoring activities in probabilistic formulations. In the last part of the paper, by relating

    damage to reliability with respect to collapse, the generalized maintenance cost for a concrete bridge

    deck was analyzed in order to verify the economic advantages offered by dynamic monitoring.

    & 2009 Elsevier Ltd. All rights reserved.

    1. Introduction

    According to the probabilistic methods used most widely for

    the assessment of safety in the structural field, reliability is

    defined as the probability of the structure attaining a limit state

    during a predetermined period of time. In many instances, this

    assessment is summarised by an ad hoc reliability index [1].

    Albeit useful at the design stage, this approach displays some

    limitations when dealing with existing structures:

    it does not take into account the additional knowledgeobtained from monitoring activities;

    it often overlooks the fact that safety deteriorates over time; it does not provide the information needed for a long-term

    evaluation of the economic convenience of restoration works.

    In particular, the latter task calls for a knowledge of the residual

    lifetime or service time of a structure, which must be a factor in

    the assessment of the overall economic utility of a strengthening

    intervention.

    New studies on reliability-based reassessment of structures

    focused on updating the probability of failure according to new

    information coming from existing structures [24]. Probability

    models may thereafter be enhanced by collecting new data

    regarding geometry, material properties, structural deterioration,loading on the structure, static and dynamic behaviour [2].

    If the degradation of reliability over time is taken into account,

    the lifetime of a structure is a random variable [5] and reliability

    can be characterised in relation to the so-called hazard function

    (the damage rate in the infinitesimal time interval), which

    assumes various forms depending on the distribution model

    adopted (Weibull, Gamma, Frechet, etc.). In this connection, a

    monitoring-oriented approach [6] is of great interest, as the

    monitoring process is able to supply useful data both to plot the

    reliability curves, defined as a function of the symptom, and to

    interpret the diagrams obtained.

    Structural monitoring, construed as a system that provides on

    request data regarding a specific change, or damage, occurring in a

    structure, can be a valid tool to fine-tune reliability estimates inthe light of the actual conditions of a structure. With structural

    monitoring systems reference is understood to devices (hardware)

    and procedures (software) used to acquire the time evolution of

    parameters that are supposed to be related to the safety condition

    of a structure (usually strains, displacements, velocities, accelera-

    tions, temperatures, forces). Today, the trend is to consider

    information coming from monitoring systems as crucial to

    decisions about retrofitting existing structures [7]. Recently,

    research focused on the possibility of obtaining more information

    on the safety condition of a structure on the base of vibration

    measurements [8,9]. The basic idea behind current dynamic

    monitoring techniques is that modal parameters are a function

    of the physical properties of the structure; therefore, changes

    ARTICLE IN PRESS

    Contents lists available at ScienceDirect

    journal homepage: www.elsevier.com/locate/ress

    Reliability Engineering and System Safety

    0951-8320/$- see front matter & 2009 Elsevier Ltd. All rights reserved.doi:10.1016/j.ress.2009.02.010

    Corresponding author.

    E-mail address: [email protected] (R. Ceravolo).

    Reliability Engineering and System Safety 94 (2009) 13311339

    http://www.sciencedirect.com/science/journal/magmahttp://localhost/var/www/apps/conversion/tmp/scratch_5/http://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.ress.2009.02.010mailto:[email protected]:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_5/dx.doi.org/10.1016/j.ress.2009.02.010http://localhost/var/www/apps/conversion/tmp/scratch_5/http://www.sciencedirect.com/science/journal/magma
  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    2/9

    in the physical properties will cause detectable changes in

    modal properties. The advantage of this approach is that a local

    measurement can provide information related to the global

    behaviour.

    In this paper, the symptom-based approach is analyzed, in

    order to evaluate its applicability to vibration-based structural

    monitoring. In the last part of the paper, also based on reliability

    with respect to the ultimate limit state (ULS) of concrete bridges, acost analysis for concrete bridge decks was performed in order to

    verify the economic advantages offered by dynamic monitoring.

    2. Symptom-based approach to safety assessment

    Let us now examine a symptomatic approach to the evaluation

    of the performance of a structure, so that reliability appreciation

    depends on measurable quantities. It is first assumed that when a

    symptom exceeds an assigned value, Sl, the structure does not

    fulfil the requirements for which it has been designed, and the

    unit is definitely in need of repair or replacement [6]. In practice,

    an excessive value of the symptom (e.g. the deflection of a bridge

    or, as in our case, its fundamental period) would result inthe structure being excluded from the monitoring program. If

    the reliability of a structure, R(t), is defined as the probability that

    the time it takes a system to reach a damage limit state associated

    to the structures lifetime, tb, is greater than a generic time t:

    Rt Ptptb, (1)

    then reliability can be rewritten as a function of the symptom

    variable, S; in this case, it is defined as the probability that a

    system, which is still able to meet the requirements for which it

    has been designed (SoSl), is active and displays a value of S

    smaller than Sb, where Sb is the value of the symptom

    corresponding to the reference limit state. Accordingly, reliability

    is defined as

    RS PSpSbjSoSl

    Z1S

    fSdS, (2)

    i.e., R(S) can be expressed by the integral of the symptoms

    distribution probability density fS. With the symptomatic ap-

    proach it is also possible to work out, for the R(S) function,

    expressions similar to those used by the time-based approach,

    that is to say for R(t); the hazard function, h(t), specifies the

    instantaneous rate of reliability deterioration during the infinite-

    simal time interval, Dt, assuming that integrity is guaranteed up

    to time t [5]:

    ht limDt!0

    PtptbotDtjtbXt

    Dt. (3)

    h(t) is connected to the reliability function, R(t), by the following

    relationship:

    Rt exp

    Zt0

    ht0dt0

    . (4)

    In a similar manner, the so-called symptom hazard function,

    h(S), is defined as the reliability deterioration rate per unit of

    increment of the symptom:

    hS limS!0

    PSpSboSDSjSbXS

    DS, (5)

    hence

    RS exp ZS

    0hS0 dS0

    !(6)

    or equivalently [5]

    hS 1

    RS

    d

    dSRS. (7)

    Iftb is the time of attainment of a damage limit state or the total

    lifetime, reliability as a function of the symptom gives the residual

    damage capacity, DD, of the structure:

    RS 1 DS DDS, (8)

    where D t(S)/tb represents the systems aging as well as the

    measure of the damage. In practical applications, symptom

    models are chosen that lead to realistic expressions for the

    residual damage capacity. For instance in structural systems,

    which are characterized by aging or failure processes, models with

    increasing hazard functions (exponential, Weibull, gamma, log-

    normal distributions, etc.), are used the most [5].

    Eq. (8) lends itself to a diagnostic use: assuming that one knows

    the evolution of reliability through the observation of a set of

    systems, the value of the symptom as observed in a given unit makes

    it possible to determine the residual lifetime of the unit itself.

    Under this approach monitoring plays a key role, in that

    reliability is no longer expressed as a function of time, but rather

    as a function of a symptom, which is a measurable quantity.

    3. Extension to structural classes

    Reliability can be described starting from a primary reliability,

    R0(S), that applies to a given type of systems (structural class)

    and can be characterised for a particular system by the introduc-

    tion of a logistic vector Li, with i 1yN, where N is the total

    number of systems to be monitored [10]. Li denotes the individual

    element of the sample, it may contain a series of specific

    parameters depending on which aspect of the system we know

    or we want to monitor.

    Each unit of the class may differ from the other elements in its

    original characteristics as well as its usage (e.g. actions ormaintenance quality). Any additional information or measure-

    ments, directly or indirectly related to the symptom S, i s a

    potential component of the logistic vector, as long as it is referred

    to the single unit. In principle this may be geometry, loads,

    environment parameters, material properties, soil, maintenance

    levels, other factors even of a binary nature.

    The Li vector appears in the formulations of system reliability

    starting from h(S), which depends on L:

    hSL hS; L, (9)

    whence, by integration and by analogy with Eq. (5), we can

    express the value of reliability, R(S,L), as a function of the

    symptom considered and the L vector:

    RS; L exp

    ZS0

    hS0; LdS0( )

    . (10)

    For the hazard function, we start from a general multiplicative

    form of the primary hazard function:

    hS; L h0SgL (11)

    in order to explore how the logistic vector, L, affects the survival

    function R(S,L). In the assumption of small changes of L, system

    reliability can be determined as [10]

    RS; LjL0 DL R0S; L0 1 DLTqg

    qLlnR0S; L0

    & '(12)

    being : R0S; L0 exp ZS

    0h0S0gL0 dS0

    ( ), (13)

    ARTICLE IN PRESS

    R. Ceravolo et al. / Reliability Engineering and System Safety 94 (2009) 133113391332

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    3/9

    where g(L) is any function satisfying the condition g(L0) 1, and

    DL is a vector containing the variations made to the parameters of

    the L vector. When g(L) is assumed to be linear Eq. (11) tends to a

    Cox proportional model [11].

    4. Interaction between monitoring and repairing costs

    In order to analyze the interaction between structural

    reliability and maintenance costs, directed to restore the initial

    reliability of the structure, in the following we shall refer to the

    ultimate limit state associated with structural failure, as related to

    a limit state of damage (DLS) directly correlated to the structures

    lifetime.

    Though a full reliability approach is a possible option, in

    accordance with previous works in the field of structural

    engineering, it is assumed that safety at the ultimate limit state,

    as expressed by index bULS F1(1RULS), where F

    1 is the

    inverse of the standard normal cumulative function [1], is

    indirectly correlated to the residual life of the structure. The

    following expression is thus used for the cost Ci of a generic

    maintenance intervention [12]:

    Ci fb0ti;Db0ti, (14)

    where Ci is the cost of the ith intervention, b0(ti) is the reliability

    of the structure at the time of the intervention performed at time

    ti, Db0(ti) is the increase in reliability caused by the intervention.

    Hence, in addition to depending on type of intervention (whose

    effect is Db0(ti)), the variability of maintenance costs is also

    affected by the health conditions of the structure at the time

    maintenance works are performed; in general, structures in good

    conditions require lesser amounts compared to rundown struc-

    tures, the reliability level attained and the type of intervention

    being the same.

    In particular, the cost function C1, used in this analysis and

    relating to a single maintenance intervention applied at time t1,

    includes a fixed part, C0, which does not depend on theimprovement achieved with the intervention, and a variable part,

    which is a function ofDb0(t1) [12]:

    C1 C0 lsDb0t1q, (15)

    where s and q are cost parameters, l is a multiplication factor that

    takes into account the reliability level b0(t1) of the structure at the

    time the intervention is performed.

    The increase Db0(t1) in reliability b0(t1) obtained with the

    maintenance intervention is designed to restore the reliability of

    the structure to the initial value of the reliability index. It is

    assumed that the value ofDb0(t1), induced by the maintenance

    intervention performed at time t1, can vary beginning from a

    threshold value that has to be assured anyway in case of

    intervention. The multiplier l is determined as follows [12]:

    l p1b0t12 p2b0t1 p3, (16)

    where p1, p2 and p3 are coefficients that vary as a function of type

    of intervention.

    Due to the maintenance intervention at time t1, the reliability

    profile is updated as follows:

    bt; t1 b0t; t0ptot1;

    b0t Dbt1; t1pt;

    ((17)

    where b0(t) is the profile of the reliability index before the

    intervention, Db(t1) is the increase in the reliability index caused

    by the maintenance intervention, and b(t,t1) is the evolution of

    reliability following the intervention applied at t1.

    Another factor to be considered in addition to the maintenancecost, is the failure risk cost Cpf1;b0

    to be evaluated from profile

    b(t,t1) as modified, with respect to the virgin index, b0(t), by the

    intervention performed at t1 [12]:

    Cpf1;b0

    cfth

    Ztht0

    Dbt; t12

    1 vtdt, (18)

    where th is the time period encompassed by the analysis of costs,

    Db(t,t1) is the deviation of b(t,t1) from the value of b0 at initial

    time t0, Db(t,t1) b0(t0)b(t,t1) if b0(t0)Xb(t,t1), otherwiseDb(t,t1) 0; cf is the coefficient reflecting the risk cost, v is the

    discount rate of money. The magnitude of cf depends on several

    factors, such as type of structure, the volume of daily traffic, the

    number of accidents, and service disruption.

    The total cost function CTOT, depending on the maintenance

    intervention time t1, is the sum of the maintenance intervention

    cost C1 (Eq. (15)) and of the risk cost Cpf1;b0:

    CTOTt1 C1t1 Cpf1;b0t1. (19)

    Let us now consider the evaluation of the economic conve-

    nience of monitoring the generic structure. If the reliability profile

    b(t) of the bridge deviates at time t1 by DDb(t1) from b0(t), i.e., the

    value that applies to the entire class of structures, the main-

    tenance interventions, selected on the basis of the values of curveb0(t), will not be able to restore b(t1) to the initial value b0(t0), and

    will only obtain a lower value: b0(t0)DDb(t1). The advantage

    offered by the monitoring process, and hence by a correct

    knowledge of the actual profile, b(t), compared to the standard

    one, b0(t), makes it possible, with an additional maintenance cost

    incurred to restore b(t1) to b0(t0), to avoid the risk cost associated

    with DDb(t1) during the time following the intervention (Fig. 1).

    The expression for the determination of the failure risk cost,

    Cpf1;b , that applies to a generic evolution b(t) (other than b0(t)) and

    to the maintenance intervention at time t1, becomes:

    Cpf1;b Cpf1;b0 cfth

    Ztht1

    DDbt12 2DDbt1Dbt; t1

    1 vtdt

    cfth

    Zt1t0

    DDbt2

    2DDbtDbt; t1

    1 vtdt, (20)

    where from the risk cost Cpf1;b0for the standard reliability profile,

    b0(t), we subtract the risk cost for the time t1th, avoided thanks to

    the monitoring and we add the additional risk cost for the time

    t0t1. DDb(t) is the deviation of the monitored reliability profile

    b(t) from the standard reliability profile b0(t). The total cost

    function CTOT, depending on the maintenance intervention time t1,

    becomes:

    CTOTt1 C1t1 Cpf1;b t1 Caddt1. (21)

    Besides the maintenance intervention cost C1, the cost of the

    additional maintenance Cadd is included in Eq. (21) and is obtained

    by the following formula (22), that is analogous to Eq. (15):

    Cadd lsDDbt1q. (22)

    Whenever relevant, also the cost of monitoring may be included

    in Eq. (21).

    5. Dynamic monitoring of bridge decks

    The application example proposed below (Fig. 2) uses simply

    rested prestressed bridge beams (95 m span, box section, fck 40

    N/mm2, fctm 3.5N/mm2, elastic modulus Ecm of concrete 35

    kN/mm2, area of the cross-section Ac 11.53 m2, and moment of

    inertia JG 23.086m4).

    The symptom that we assume to monitor over time, through

    customary experimental modal analysis procedures is the funda-mental period, T, of this structural class, whose variation is

    ARTICLE IN PRESS

    R. Ceravolo et al. / Reliability Engineering and System Safety 94 (2009) 13311339 1333

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    4/9

    associated with the decrease in stiffness. Due to the purely

    methodological value of this example, the symptoms evolution

    is not measured but calculated according to an analytical

    damage model. The first cracking moment of the deck, Mcr,

    is 53MNm, as determined according to the following

    formula [13]:

    Mcr fctmWi, (23)

    where fctm 3.5N/mm2 is the average tensile strength of concrete

    and Wi is the section modulus of the uncracked section at the

    lower chord.

    Having defined a log-normal statistical distribution of the live

    load q on the bridges referred to a 1-year period (mean value:

    45 kN/m, variation coefficient: 0.15), the analysis is performed for

    each time step and each load level envisaged [1,13,14]. The spatial

    distribution of the loads was performed according to the rules set

    forth in the European standards [13,14]. Another option would be

    to turn to bimodal load distributions [15], but this is far beyond

    the scope of the present paper.

    The damage model considered is based on the elastic theory of

    damage [16], so that the deterioration process in the bridges,

    triggered when the damage threshold envisaged was exceeded,translates into a reduction in bending stiffness, EI, according to the

    following expression [17]:

    EI EI01 d, (24)

    where EI0 is the stiffness of the uncraked section, and d is the

    damage parameter. The isotropic damage parameter d can be

    physically interpreted as the ratio of damaged surface area,

    corrected by stress concentration effects and interactions, over

    total surface area at a local material element [18].

    The deformation energy, corresponding to the first cracking

    moment, Mcr, is assumed as the threshold value, x0, beyond which

    the damage mechanism is triggered in the beam.

    For a given time step, each statistical value of the load q

    corresponds to an accumulated deformation energy, x, and a certain

    size of the crack zone in the beam astride its midspan. The damage

    to the deck, reflected by parameter d, affects only the cracked zone

    and propagates over time according to the elastic model: at the n+1

    interaction, the elastic deformation energy, xn1n1, a function ofthe state of strain, en+1, is determined; we get the damage parameter,dn+1, and the damage threshold, rn+1, as given below [16]:

    dn1 dn if xn1orn;

    1 1 Ax0=xn1 A expBx0 xn1;(rn1 maxrn;xn1. (25)

    ARTICLE IN PRESS

    (t)

    (t)

    0(t)

    Additional failure safety

    deterioration from monitoring

    1

    1

    Failure safety deterioration for the standard

    reliability profile

    Failure safety advantage from

    monitoring

    Standard reliability

    profile

    Reliablity profile obtained by

    structure monitoring

    t1t

    Fig. 1. Advantage offered by structural monitoring when the reliability profile is lower than the standard reliability profile. t1 is the time of the maintenance intervention.

    1270 cm

    40

    335250

    45

    510

    70

    45335

    Fig. 2. Bridge deck section. Measurements are in centimeters.

    R. Ceravolo et al. / Reliability Engineering and System Safety 94 (2009) 133113391334

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    5/9

    In Eq. (25), the deformation energy, xn+1, is compared with the

    limit, rn, and the damage parameter, dn+1, is maintained the same as

    in the previous step if the accumulated energy, xn+1, does not exceed

    the limit rn; conversely, dn+1 is defined with Eq. (25) ifxn+1 exceeds

    rn. In Eq. (25), A and B stand for the growth coefficients of the

    damage law, which are 0.68 and 1.41, respectively, for high strength

    concrete [18]. The fundamental period was calculated by averaging

    over simulated outcomes referred to a single year, as resulting fromMonte Carlo simulations based on the distribution assumed for

    loads.

    It should be noted that with the build-up of the damage

    resulting in the decrease in stiffness, EI, the fundamental period T

    of the structure increases asymptotically up to a value Tf of about

    1.25 s, with a 14% increment over the initial value of the period,

    T(t 0), which was 1.09 s.

    The residual damage capacity of the system, or its reliability as

    a function of the symptom observed, R(S), is given by [6]

    RS DDS 1 tS

    tb. (26)

    In this application it was assumed for the bridges lifetime

    tb 100 years.In evaluating the reliability of existing structures one cannot

    rely on monitoring data covering the entire life of a structure, on

    the other hand it has been ascertained that a few initial data

    regarding the symptom observed over time are sufficient to

    identify the underlying trend evidenced by it. The symptom-based

    approach makes it possible to choose, from among different

    variation curves of the symptom over time, the one that best

    reflects the trend observed so as to obtain a tool for the evaluation

    of the current and future conditions of the system.

    By way of exemplification, the evolution over time of the

    symptom/fundamental period can be approximated with a

    lifetime distribution model [5] (Fig. 3a). For instance, in this case

    a possible option would be a Weibull model:

    S=St0 1 1a ln1 t=tb

    1=g (27)

    with the coefficients a 4.6 and g 6.2. Correspondingly thereliability (Fig. 3b) would become:

    RS expfSSt0 1a

    gg, (28)

    whose associated hazard function h(S) is monotone increasing

    (g41) with the symptom (Fig. 4) [5]. Apparently, in this example,the Weibull and the Frechet models tend to overestimate

    reliability in the short/medium period, while they are conserva-

    tive when the bridges service time is approaching tb. Obviously,

    the selection of a specific model will depend on the monitoring

    experiences performed on different structural types.

    Knowing the current value ofS, Eq. (28) supplies an evaluation

    of the current and future conditions of the structural class in

    terms of residual lifetime or primary reliability.

    Given the primary reliability function, R0(S,L0), valid for a

    family of structures of the same type, by monitoring a single unit

    in the class it is possible to calibrate with greater accuracy theestimate of its residual damage capacity, R(S,L).

    If monitoring results reveal an evolution of the symptom faster,

    or slower, than the standard rate assumed for the structural

    family, the estimate for the specific structure in question can be

    modified through Eq. (12), where R0(S,L0) is the survival function

    for the standard structure, and R(S,L)L0+DL is the survival function

    for a specific structure characterised by increment DL of the basic

    logistic vector, L0.

    If R0(S,L0) is made to coincide with R0(S) and the measured

    fundamental period Tm is the only monitored quantity to be

    inserted in the logistic vector, the following form may be assumed

    for Eq. (11):

    hS; L h0SgL h0SL h0ST

    mT0 . (29)

    ARTICLE IN PRESS

    1.161.14

    1.12

    1.1

    1.08

    1.06

    1.04

    1.02

    T/T(t=0)

    0

    1

    0.02 0.04 0.06 0.08 0.1

    t / tb

    Elastic theory of damage

    Exponential type distribution model

    Weibull distribution model

    Frechet distribution model

    1

    0.99

    0.98

    0.97

    0.96

    0.95

    0.94

    0.93

    0.92

    0.91

    0.9

    R

    1 1.05 1.1 1.15

    T / T (t =0)

    Fig. 3. (a) Symptom evolution by the elastic theory of damage and by statistical models. (b) Damage limit state reliability as a function of the symptom.

    12

    10

    8

    6

    4

    2

    0

    Sym

    to

    mhazard

    func

    tion

    1.08 1.1 1.12 1.14 1.16 1.18 1.2 1.22 1.24 1.26

    Fundamental period T

    Fig. 4. Symptom hazard function h(S) resulting from a Weibull model. In this case

    the symptom is the bridge decks fundamental period T.

    R. Ceravolo et al. / Reliability Engineering and System Safety 94 (2009) 13311339 1335

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    6/9

    In other words, hazard function h(S,L) is assumed to be modified

    proportionally to the measured symptom Tm referred to its

    primary value, T0. For generalitys sake, in practical applications

    Tm and T0 may be conveniently referred to their initial values.Correspondingly Eq. (12) becomes

    RS; LjL0 DL R0S 1 DT

    T0lnR0S

    & ', (30)

    where a positive deviation in the symptom, DT TmT0, indicates

    a reduction in reliability. The evolution of the fundamental period

    of the structure in time is represented in Fig. 5a, while the graphs

    of RDLS(t), corresponding to different evolutions of natural period,

    are shown in Fig. 5b.

    6. Interaction between reliability and costs

    A cost analysis, according to Section 4, has been applied to thebridge deck. For the sake of simplicity, here it is assumed that

    reliability with respect to failure (RULS) is indirectly related to

    reliability with respect to damage (RDLS), which governs the

    structures residual lifetime.

    The reliability index profiles, bDLS (Fig. 5d) have been obtainedfrom RDLS through the following relationship:

    bDLSt F11 RDLSt. (31)

    Likewise the graphs ofbULS(t) (Fig. 5c), have been obtained from

    the reliability RULS. Curves in Fig. 5c refer to different values ofDT/

    T0 (0.1, 0.2, 0.5, respectively), virtually found with monitoring.

    The evolution of the total cost, including the maintenance cost

    and the risk cost, has been obtained through the formulas (20)

    and (21) as a function of the intervention time (Fig. 6). The

    adopted values for the cost parameters associated to the chosen

    type of intervention are indicated in Table 1.

    If the reliability profile bULS(t) is lower than the standard one,

    bULS,0(t), the monitoring process results to be advantageous from

    the economic standpoint, as long as the risk cost avoided exceedsthe additional maintenance cost.

    ARTICLE IN PRESS

    1.16

    1.14

    1.12

    1.1

    1.08

    1.06

    1.04

    1.02

    1

    T/T(t=

    0)

    0 5 10 15 20

    Time (years)

    L = T/T0= 0.5

    Elastic theory of damage

    Monitored symptom profiles L >0

    Monitored symptom profiles L

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    7/9

    Then the graph ofbULS(t) has been approximated through the

    function [12]:

    bULSt bULSt0 at t00:5, (32)

    where the degradation parameter a0 is assumed to be 0.109 for thestandard profile bULS,0 (Fig.1) and bULS(t0) is assumed to be 5.5. If a

    uniform probability distribution is assumed for a parameter [12](see Fig. 7b), different variation coefficient, c

    a, for the statistical

    distribution produce the graphs shown in Fig. 7a: the higher the

    variation coefficient the more advantageous the effects ofstructural monitoring. Correspondingly the maintenance inter-

    ventions result to be slightly anticipated with monitoring.

    7. Modal testing: sensitivity of different parameters

    The field of structural identification now offers a vast range of

    effective techniques. In the civil engineering field, of special

    interest are methods which do not require a prior knowledge of

    the dynamic input and are able to take advantage of the natural

    excitation to which a structure is subjected, so as to enable the

    behaviour of the structure to be monitored in operating condi-

    tions [19]. In recent years, time domain techniques have been

    used rather successfully [20], thanks to the great spectralresolution offered and to their modal uncoupling capability.

    The situation is more critical for damping estimation, since this

    parameter, having no significant effects on frequency, primarily

    affects the modulation of modal signals and, in unknown input

    conditions, becomes latent information. In actual fact, the

    accuracy in damping estimation afforded by current output-

    only methods is not very high [21]. These considerations

    prompted some proposals for timefrequency methods, which

    are able to handle non-stationary excitation typical of bridges and

    other civil structures [22], but, at the same time, bring about

    complexity and computational cost. We conclude that, while

    modal frequencies may be evaluated efficiently through standard

    output-only identification procedures, damping monitoring in

    civil structures still requires the excitation to be measured and

    this may prove costly. This notwithstanding, in the following we

    present an example in which both frequency and damping havebeen ideally monitored.

    The numerical application described below is about reinforced

    concrete bridge piers (H 5 m, section diameter + 1.1 m,

    fck 40 N/mm2, concentrated mass at the top of the

    pier 410,000 kg, geometric reinforcement ratio rl 2.41% andhorizontal design load Hd 868 kN, initial cracking moment for

    the section Mcr 1.13 MN m).

    The symptoms that we monitor over time are the fundamental

    period of the piers, whose variation is associated with the

    decrease in stiffness, and an equivalent viscous damping. Damp-

    ing is obtained from forced vibrations (vibrodyne).

    The difference with the model used in Section 5 concerns

    essentially in damping: the results obtained on reinforced

    concrete structures, in fact, have shown that, in this material,stress intensity, i.e., cracking state, has a decisive influence on

    ARTICLE IN PRESS

    3000

    2500

    2000

    1500

    1000

    500

    0 10 20 30 40 50

    Time of the maintenance intervention (years)

    0.2

    0.5

    L = T /T0

    = 0.1

    Generalized cost without structure monitoring

    Generalized cost with structure monitoring

    Totalcost

    /m2

    Fig. 6. Generalized cost as a function of the time of the maintenance intervention:

    curves for different evolutions of the natural period (discount rate n 2%)U

    Table 1

    Bridge deck: cost parameters associated to the chosen type of intervention [9].

    Type of intervention Fiber-reinforced polymer

    attaching

    Fixed part of the intervention cost, C0 400$/m2

    Time period encompassed by the analysis of costs, th 50 years

    Cost parameter, s 230

    Cost parameter, q 2

    The coefficient reflecting the risk cost, cf 4000$/m2

    The discount rate of money, v 2%

    Parameters associated with parabolic function for

    multiplier, l

    p1 0.25

    l p1b2+p2b+p3 p2 2.0

    p3 5

    3000

    2500

    2000

    1500

    1000

    500

    0

    0

    10 20 30 40 50

    Time of the maintenance intervention (years)

    c =0.35

    c =0.35

    c

    =0.45

    c =0.45

    c =0.55

    c =0.55

    1

    0.5

    0 0.005 0.024 0.043 0.175 0.194 0.213

    CDF

    To

    talcost

    /m2

    Fig. 7. Generalized cost as a function of the time of the maintenance intervention

    (mean value): (a) curves for different values of the variation coefficient ofa, ca and(b) cumulative distribution functions for different values of the variation

    coefficient ca.

    R. Ceravolo et al. / Reliability Engineering and System Safety 94 (2009) 13311339 1337

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    8/9

    equivalent damping. The value of this parameter is seen to

    increase with increasing stress level until the structural element is

    fully cracked; after cracking, damping begins to decrease [23]. In

    the application described below, the evolution of damping for

    purposes of reliability assessment is determined with reference to

    the conditions that precede the fully cracked state. To this end, a

    hysteretic RambergOsgood mechanical model has been adopted

    for the pier [24].

    The evolution over time of the equivalent stiffness, K, is worked

    out from the fundamental period via the elastic damage model

    reported above [17], and therefore the RambergOsgood model is

    updated on a yearly basis. By exciting the structure by means of a

    vibrodyne, for each step of the analysis it is possible to quantify

    the corresponding equivalent damping, xeq, according to thefollowing formula [24]:

    xeqt 2

    p1

    2

    g 1

    1

    Fvib=Dvibt

    Kt

    , (33)

    where Fvib is the dynamic load ideally generated by the vibrodyne,

    and Dvib is the ensuing displacement observed in the structure.

    If the pier is excited yearly with a vibrodyne calibrated at a

    constant value (Fvib 300kN), the evolution over time of the

    equivalent damping is determined from Eq. (33), in the assump-

    tion that the measured horizontal displacement decuples asymp-

    totically its initial value and for g 2. As g influences strongly thevariation of damping over time, in practice this parameter should

    be determined on a preliminary basis with sufficient accuracy. The

    relative variation of damping over time, as plotted in Fig. 8, clearlyshow a potentially increased sensibility of the reliability assess-

    ment procedure when also damping is monitored. The reason for

    this improvement is that, while a detectable change in the

    fundamental period is usually restricted to the first years of

    service, the damping parameter continues to increase slowly and

    consistently with the bridge effective age.

    The reliability of the monitored bridge piers, as a function of

    small variations of the logistic vector, L, is worked out from Eq. (12).

    In this case, the logistic vector, L, reflects both parameters monitored,

    i.e., fundamental period, T, and equivalent damping, xeq;DLTcontains

    the deviations of these parameters over time, Tm and xeq,m, from

    their primary values, T0 and xeq,0 and Eq. (30) becomes:

    RS; LjL0 DL R0S 1 DT

    T0;Dx

    eqxeq;0

    & ' p1p2

    !lnR0S

    ( ), (34)

    where qg/qL reduces to a weight vector (p1,p2)T to be associated

    with the two symptoms and the accuracy afforded in their

    evaluation.

    8. Conclusions

    This paper addresses the problem of the probabilistic assess-ment of the reliability of civil structures through a symptomatic

    approach, which is able to create an appropriate theoretical

    framework for taking into account, in safety checks, periodic or

    continuous monitoring activities. In particular, it lends itself to the

    use of dynamic parameters (frequencies, modal shapes and

    damping), identified either through non-destructive tests per-

    formed on existing structures or through experimental modal

    analyses conducted on structures set up to this end, for the

    estimate of the residual lifetime of a construction. By relating

    damage to reliability with respect to collapse, the generalized

    maintenance cost was also analyzed in order to verify the

    economic advantages offered by monitoring.

    Simulated applications to bridge structures, subjected to

    periodic monitoring, have been illustrated, in which two symp-toms were considered: the reduction in stiffness and the increase

    in an equivalent viscous damping. The examples showed that the

    outcome of dynamic monitoring systems in bridge structures

    might be conditioned by the availability of accurate damping

    measurements, which requires ad hoc excitation. While damp-

    ing monitoring from forced vibrations may prove very costly, it is

    also true that advances in output-only identification techniques

    are expected for the next years.

    In actual practice measurements are noisy and affected by

    different factors, whose relative importance varies with the

    structural class and the monitoring system. For instance modal

    quantities are known to be strongly affected by thermal fluctua-

    tions. A future development of this study will consist of analyzing a

    few monitoring systems by expressing measurement uncertainty.

    References

    [1] Nowak AS, Collins KR. Reliability of structures. McGraw-Hill InternationalEditions; 2000.

    [2] Diamantidis D, et al. Probabilistic assessment of existing structures. JointCommittee on Structural Safety (JCSS), RILEM Publications; 2001.

    [3] Faber MH, Srensen JD. Indicators for inspection and maintenance planning ofconcrete structures. Structural Safety 2002;24(4):37796.

    [4] Straub D, Faber MH. Risk based inspection planning for structural systems.Structural Safety 2005;27(4):33555.

    [5] Lawless JF. Statistical models and methods for li fetime data. New York: Wiley;1982.

    [6] Natke HG, Cempel C. Model-aided diagnosis of mechanical systems.Germany: Springer; 1997.

    [7] Boller C, Chang FK, Fujino Y, editors. Encyclopedia of structural healthmonitoring. Chichester, UK: Wiley; 2009.

    [8] Maeck J, Abdel Wahaba M, Peeters B, De Roeck G, De Visscherb J, De WildeWP, et al. Damage identification in reinforced concrete structures bydynamic stiffness determination. Engineering Structures 2000;22(10):133949.

    [9] Sohn H, Farrar CR, Hemez FM, Shunk DD, Stinemates DV, Nadler BR. A reviewof structural health monitoring literature: 19962001. Los Alamos NationalLaboratory Report, LA-13976-MS, 2003.

    [10] Cempel C, Natke HG, Yao JTP. Symptom reliability and hazard for systemscondition monitoring. Mechanical Systems and Signal Processing2000;14(3):495505.

    [11] Cox DR. Regression model and life tables. Journal of the Royal StatisticalSociety B 1972;34:187220.

    [12] Kong JS, Frangopol DM. Costreliability interaction in life-cycle costoptimization of deteriorating structures. Journal of Structural Engineering2004;130(11):170412.

    [13] EN 1992-2:2005 Eurocode 2: Design of concrete structuresPart 2: concretebridgesdesign and detailing rules.

    [14] EN 1991-2:2003 Eurocode 1: Actions on structuresPart 2: traffic loads onbridges.

    [15] Mei G, Qin Q, Lin DJ. Bimodal renewal processes models of highway vehicleloads. Reliability Engineering & System Safety 2004;83:3339.

    ARTICLE IN PRESS

    1.6

    1.5

    1.4

    1.3

    1.2

    1.1

    1

    eq

    /

    eqt

    =0

    0 5 10 15 20

    Time (years)

    eq

    eq, 0=0.3

    Elastic theory of damage

    Monitored symptom profiles L >0

    Monitored symptom profiles L

  • 7/29/2019 2012-12-26-Symptom_based Reliability and Generalizedre Pairing Cost

    9/9

    [16] Ju JW, Monteiro PJM, Rashed AI. Continuum damage of cement paste andmortar as affected by porosity and sand concentration. Journal of EngineeringMechanics 1989;115(1):10530.

    [17] DiPasquale E, Ju JW, Askar A, Cakmak AS. Relation between global damageindices and local stiffness degradation. Journal of Structural Engineering1990;116(5):144056.

    [18] Lemaitre J. How to use damage mechanics. Nuclear Engineering Design1984;80:23345.

    [19] Peeters B, DeRoeck G. Reference-based stochastic subspace identification for

    output-only modal analysis. Mechanical Systems and Signal Processing 1999;13:85578.[20] Cunha A, Caetano E, editors. System identification and modal updating. In:

    Proceedings of the experimental vibration analysis for civil engineeringstructures (EVACES07) conference, FEUP, Porto, 2007 [Chapter 6].

    [21] Brincker R, De Stefano A, Piombo B. Ambient data to analyse thedynamic behaviour of bridges: a first comparison between differenttechniques. In: Proceedings of the 14th international modal analysisconference, Society of Experimental Mechanics, Bethel, CT, USA; 1996.p. 47782.

    [22] Ceravolo R. Use of instantaneous estimators for the evaluation ofstructural damping. Journal of Sound and Vibration 2004;274(12):385401.

    [23] Chowdhury SH. Damping characteristics of reinforced and partially pre-

    stressed concrete beams. PhD thesis, Faculty of Engineering, GriffithUniversity, Australia, 1999.[24] Otani S. Hysteresis models of reinforced concrete for earthquake response

    analysis. Journal of Faculty of Engineering, University of Tokyo 1981;36(2):40741.

    ARTICLE IN PRESS

    R. Ceravolo et al. / Reliability Engineering and System Safety 94 (2009) 13311339 1339