Sie sind auf Seite 1von 10

50

Damage indices and damage measures


E Cosenza and G Manfredi Universit` a di Napoli Federico II, Italy

Summary
This article reviews the ground motion parameters that can be assumed as structural and non-structural damage measures. Measures of seismic damage potential based on ground motion records are rst described, followed by a discussion of the damage measures relating to simple (linear) and more complex (non-linear) structural responses. The second section reviews the measures of damage Prog. Struct. Engng Mater. 2000; 2: 5059 phenomena which govern structural degradation and/or collapse, including experimental results and analytical models. The relationship between earthquake characteristics and type and level of damage on the seismic response of structures is examined, and data from different well-known destructive earthquakes are given.

Earthquake-resistant design versus damage indices and measures


In recent years, the development of structural design criteria for new structures[1**] and the renewed importance of the assessment of seismic vulnerability of existing under-designed buildings[2] have broadened the objectives of seismic design. While safety against collapse is still the main goal, performance in terms of functionality and economy have assumed a central role in the design criteria[3]. Hence, a great effort has been made to improve the current earthquake-resistant design methods in order not only to avoid collapse under a destructive earthquake, but also to limit the damage under moderate earthquakes. Furthermore, the new design philosophy is tending to multi-level probabilistic structural performance criteria, replacing completely the simple force strength approach[4**,5*]. However, the implementation of all these new concepts requires the definition of a quantative damage index and measures.

A key issue in the performance-based design is a reliable assessment of seismic damage potential. If the expected intensity of the earthquake is greatly overestimated, the cost of new construction and seismic rehabilitation of existing structures could be excessive. On the other hand, if the intensity is seriously underestimated, the results may be heavy damage and loss of life. To this end, a reliable definition of seismic intensity has to relate to the effect of damage on structural behaviour in order to assess the potential seismic hazard and to classify the seismic input. Different reports have been published recently on the use of a damage index and damage measures in earthquake engineering. They aim to clarify the different approach methodologies[6**8] and to detail different proposed formulations[911]. This review will focus mainly on the crosscorrelation between ground motion parameters used as damage potential measures and structural and nonstructural damage measures.

Abbreviations EPH = elastoplastic with hardening EPP = elasto-perfectly-plastic ESDOF = elastic single-degree-offreedom EPS = elastoplastic with softening PGA = peak ground acceleration PGD = peak ground displacement PGV = peak ground velocity RC = reinforced concrete

RMSA = root mean square acceleration RMSD = root mean square displacement RMSV = root mean square velocity SDOF = single-degree-of-freedom

Terminology Ae = elastic spectral acceleration ag = ground motion acceleration Ap = plastic spectral acceleration b = low-cycle fatigue parameter = Park & Ang parameter C = amplication factor D = damage index d = displacement dg = ground motion displacement De = elastic spectral displacement Eh = hysteretic energy

EI Fy IA ID IS m n neq o

= = = = = = = = = =

input energy yielding force Arias intensity damage factor Housner spectral intensity average of the plastic cycles, mass ductility number of plastic cycles number of equivalent cycles intensity of zero crossing

PD = R = T = T1,T2 = tB tD tE Ve vg = = = = =

Saragoni index reduction factor elastic period/time periods of Newmark & Hall representation bracketed duration effective duration record duration elastic spectral velocity ground motion velocity

Copyright 2000 John Wiley & Sons, Ltd

Prog. Struct. Engng Mater. 2000; 2: 5059

DAMAGE INDICES AND DAMAGE MEASURES Ground motion parameters and damage measures
In this section, damage measures based on data from earthquake records alone, without any structural response data, will be described. The simplest way to characterize the ground motions is in terms of peak or integral parameters. Table 1 reports the values of ground motion parameters for some destructive earthquakes.
Peak parameters

51

shown by the excellent agreement of their correlation studies with the observed damage in the Chilean earthquake of 1985[16]. Cosenza & Manfredi have proposed a damage factor, ID, that is related to the number of plastic cycles, n, and therefore, to the energy content of the earthquake[28]: (4) All the integral measures depend upon the duration of the earthquake which is a measure that cannot be predicted with any certainty. Furthermore, experiences from different earthquakes have confirmed that the duration of ground motion largely influences the level of structural damage. Records with large acceleration and spectral values produce slight damage if the duration is short (eg the Ancona earthquake in 1972), whereas records with low acceleration and long duration can be very destructive (eg the Mexico earthquake in 1985)[17*]. The earthquake duration, tE, is an unpredictable measure, as already mentioned, hence, defining earthquake duration in relation to the main energy content is a problem. In regard to this, Trifunac & Brady[18] have defined the effective duration, tD, as the time elapsed between 5% and 95 % of the RMSA; Kawashima & Aizawa[19] have introduced the bracketed duration, tB, which they define as elapsed time between the first and last acceleration being greater than a percentage of PGA. Trifunac & Novikova[20] have recently proposed a more refined determination of tD as the sum of the record intervals with a total amount of RMSA greater than 90%. Such a definition is more effective for an earthquake characterized by a chain of two or more events, such as the Campano-Lucano earthquake in 1980.

The peak parameters are peak ground acceleration (PGA), velocity (PGV) and displacement (PGD), and peak motion ratios PGV/PGA and PGD/PGV. The PGA is a basic measure of earthquake potential but is not totally reliable. Examination of recorded seismic events has shown that earthquakes with a very large PGA could not produce appreciable structural damage, while earthquakes with a very low PGA could produce an unexpectedly high level of destruction. Instead, the PGV seems to be a more representative measure of earthquake intensity as it is directly connected with energy demand[12]. The peak motion ratio PGV/PGA is indicated by different authors as being a measure of destructiveness, for instance in refs[13,14] the ground motions with larger damage potential show higher PGV/PGA values. Furthermore, this peak ratio can be related to energy demand as discussed below.
Integral parameters

The integral parameters are the root mean square acceleration (RMSA), velocity (RMSV) and displacement (RMSD) defined as: (1) where x(t) is either the ground acceleration ag(t), velocity vg(t) or displacement dg(t) and tE is the total duration of the earthquake. These integral parameters are obviously much more effective for measuring the energy content of a seismic event. Two measures of earthquake destructiveness based on the RMSA are the Arias intensity, IA[15], defined as follows: (2) and the Saragoni Factor, PD[16], given by: P D= I A/ o2 (3)

Linear response behaviour and damage measures


In this section, measures of damage potential of an earthquake based on parameters relating to the response of a linear elastic single-degree-of-freedom (ESDOF) system are described. Acceleration, velocity and displacement response spectra represent the response of ESDOF systems. The fundamental period, T, and the damping ratio are the only parameters needed to characterize dynamically the ESDOF system. However, the evaluation of these parameters gives rise to some uncertainty in the case of reinforced concrete (RC) elements because of cracking, and steel structures, as well, because of joint slippages. Elastic spectral shape and values can be assumed to be basic measures of earthquake potential. A simple and comprehensive representation of elastic spectra is given by NewmarkHall[21]. The representation is characterized by fixed values of pseudo-acceleration Ae(T) = CaPGA, of pseudo-velocity Ve(T) = CvPGV
Prog. Struct. Engng Mater. 2000; 2: 5059

where the intensity of zero crossing o is the number of zero crossings of the record in the time unit. The Arias intensity can be related to energy content[11], whereas Araya & Saragoni have shown that PD is a more effective measure of earthquake destructiveness (in terms of expected ductility) as
Copyright 2000 John Wiley & Sons, Ltd

52
Table 1 Ground motion parameters for destructive earthquakes
PGA Earthquake Nahanni 1985 Kobe 1995 Chile 1985 Ancona 1972 Friuli 1976 Bucharest 1977 Mexico 1985 CampanoLucano 1980 Record S1 L JMA NS Llolleo N Rocca NS Tolmezzo WE Incerc NS SCT EW Calitri WE (cm/s2) 1080.5 817.8 639.5 538.1 429.3 315.2 192.3 167.9 156.0 PGV (cm/s) 46.2 92.0 41.1 10.9 41.3 32.6 69.0 61.8 20.9 PGD (cm) 10.4 71.8 14.2 5.5 8.2 4.6 18.1 21.9 19.0

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS

PGV/PGA RMSA (s10-2) 4.3 11.2 6.4 2.0 9.6 10.3 35.9 36.8 13.4 (cm/s2) 113.0 88.5 96.3 47.3 112.0 43.0 31.5 29.3 29.6

tD (s) 7.92 8.34 35.68 2.85 10.52 4.93 15.58 38.82 47.17

IA (cm/s) 462.5 838.4 1520.8 67.8 446.2 119.9 71.4 243.8 134.1

o (1/s) 16.93 4.80 9.50 99.03 4.75 5.08 4.36 1.13 4.16

PD (cms) 1.61 36.45 16.85 0.01 19.75 4.66 3.75 189.81 7.77

ID

5.50 6.91 35.84 6.94 15.35 7.25 3.66 14.55 17.85

Montenegro 1979 Petrovac NS

and of displacement De(T) = CdPGD, in a range of low (with T < T1), medium (T1 < T < T2) and long periods (T > T2), respectively. PGA, PGV and PGD are the peak values of acceleration, velocity and displacement respectively, while Ca, Cv and Cd are the amplification factors. The fundamental periods, T1 and T2, govern the shape of the spectrum. These periods and the amplification factors can be evaluated using the maximum likelihood method from the actual spectrum. For rigid structures (T < 0.03 s), the spectral acceleration is equal to PGA, while for very low periods (0.04 < T < 0.4 T1) it varies with a linear law[22]. An example of the NewmarkHall spectral representation is given in Fig. 1. The NewmarkHall representation is not effective for all types of earthquake (ie for spectra with multiple-peaks or high-amplification periods). However, it allows the spectral shape to be described easily.

The maximum pseudo-acceleration, AeM, is a measure of the maximum strength demand of the earthquake; in fact it is proportional to the maximum seismic force acting on the structure. The maximum pseudo-velocity, VeM, is a measure of the kinetic energy of the earthquake and it is related to the peak energy demand. The Housner spectrum intensity, IS[23], defined as the integral of the pseudo-velocity spectrum in the range 0.12.5 s can be interpreted as an integral measure of the energy demand. Kappos[24] suggested a modification of the period range for better scaling criteria. The values of linear response parameters for some destructive earthquakes are given in Table 2. The maximum displacement, DeM, is a measure of the maximum displacement demand and can be associated with the maximum inter-storey drift[25] which controls structural and non-structural damage

a (m/s2)

T (s) Fig. 1 A NewmarkHall spectral representation


Copyright 2000 John Wiley & Sons, Ltd Prog. Struct. Engng Mater. 2000; 2: 5059

DAMAGE INDICES AND DAMAGE MEASURES


Table 2 Response parameters for destructive earthquakes Earthquake Record Ca Cv
Cd T1 (s) Nahanni 1985 Kobe 1995 Chile 1985 Ancona 1972 Montenegro 1979 Friuli 1976 Bucharest 1977 Mexico 1985 Campano-Lucano 1980 S1 L JMA NS Llolleo N Rocca NS Petrovac NS Tolmezzo WE Incerc NS SCT EW Calitri WE 2.2 2.3 2.4 2.6 3.0 2.7 2.7 3.1 2.9 1.6 2.8 2.7 2.6 3.0 2.2 2.3 3.4 1.8 0.4 0.6 1.3 0.6 3.0 1.5 2.2 3.8 1.9 0.200 0.875 0.450 0.125 0.625 0.525 1.700 2.500 0.700 T2 (s) 0.350 0.925 1.050 0.150 0.650 0.550 1.725 2.525 1.475 IS (cm) 154.7 416.9 223.8 29.5 186.8 92.8 208.9 275.2 100.6 EImax (cm2/s2) 10427 55902 67976 618 33623 7744 23621 228229 9876 Ehmax

53

(cm2/s2) 4611 31675 28600 409 15119 3053 7026 65486 3913

due to lateral displacements in multi-storey buildings. On average, the amplification factors assume values equal to 2.5[22] even though the magnitude, the epicentral distance and the site conditions have a large influence on them. Assuming constant values for the amplification factors, ie not dependent on the earthquake, PGA, PGV and PGD can be assumed to be damage measures instead of the corresponding spectral peak values. As far as the fundamental periods are concerned, high values of T1 represent a great range of structures subjected to the maximum strength demand and, as a result, a large amount of damage; the range T1, T2 is representative of the number of structures subjected to peak energy demand. Ref.[26] proposes that T1 and T2 could be expressed as proportional functions of PGV/PGA and PGD/PGV, respectively: therefore, large values of PGV/PGA correspond to higher values of T1.

Non-linear response behaviour and damage measures


In this section, measures of the damage potential of an earthquake based on parameters related to the nonlinear response of a SDOF system are described. The non-linear response of a SDOF system depends on the hysteretic rule used in the analysis, as discussed in the following. The choice of hysteretic rule is related to the characteristics of the modelled structure. Inelastic acceleration, velocity and displacement spectra basically represent the response of non-linear systems. The inelastic ductility spectrum can be constructed either in an exact way by implementing a nonlinear step-by-step analysis or by an approximate procedure introducing the reduction factor, R, which represents the ratio of the elastic pseudo-acceleration spectrum, Ae(T), to the inelastic, Ap(T), for fixed values of elastic period, T, and inelastic ductility = u/y. By knowing the elastic spectrum and introducing a function that allows the determination of R from the supplied ductility, , it is possible to obtain the inelastic spectrum directly, without carrying out any nonlinear
Copyright 2000 John Wiley & Sons, Ltd

analysis. The definition of consistent expressions of R is still an open question and different formulations have been suggested in the literature[27]. The plastic acceleration Ap(T,) represents for a fixed value of the supplied ductility a measure of the strength demand of the ground motion. As will be discussed in the following, this measure is representative of the behaviour of structures that collapse for maximum plastic displacement, independent of the amount of dissipated energy. On the other hand, the cyclic collapse of structures that show a degenerative behaviour and/or cumulative damage is dependent on the amount of hysteretic energy, Eh. The demand of Eh is an effective measure of the damage potential of the earthquake even though a correlation with the ground motion parameters is not simple. The energy dissipation is due to plastic cycles in the structural response. The assessment of the number of plastic cycles, n, and of the average value of their amplitude, m, allows evaluation of the equivalent number of cycles, neq[9,29]. This parameter is a measure of the distribution of amplitude cycles and indicates the number of cycles at the maximum plastic displacement (uy) which the structure has to develop in order to dissipate the total amount of hysteretic energy, Eh. Hence, values of neq close to 1 indicate the presence of one large plastic cycle and are typical of an impulsive earthquake such as the ground motion of Bucharest NS, whereas high values of neq refer to a large number of plastic cycles and are typical of long duration earthquakes as seen in the ground motion records for the Chilean earthquake of 1985. Rodriguez[30] gave a better definion of the parameter neq, referred to as a measure of damage potential. The latter is defined as the number of complete cycles of the equivalent SDOF system with a maximum displacement equal to the peak acceptable roof displacement so as to absorb the total hysteretic energy. The assessment of Eh is performed generally by knowing the input energy EI and the ratio EI/Eh given
Prog. Struct. Engng Mater. 2000; 2: 5059

54
as a function of ductility[31,32]. The input energy can be defined (in the relative formulation) as: (5) where v(t) is the relative velocity and m is the mass of the structure. The relative and the absolute formulations provide very similar results in the period range which is of practical interest[17]. The input energy from a ground motion depends mainly on the elastic period of the structure and on the seismic record, whereas it is slightly influenced by the viscous damping and by the characteristics of the plastic response[31]. For these reasons, the input energy has been assumed as a measure of the energy content of the earthquake[17]. An early relation was suggested by Housner and is valid for the undamped system[33]. It relates the peak value of the input energy, EImax, to the square of the spectral peak velocity. Another basic relationship was proposed by Akiyama[34] which relates the peak input energy to the predominant period of ground motion assumed to be similar to T1. The study of various earthquake records indicates that the peak input energy largely depends on the global work of the earthquake and therefore an integral measure must be included in a reliable formulation for the assessment of EImax[17,26] which relates the effective duration tD to the input energy, and thus tD becomes an effective measure of the intensity of the ground motion. Different formulations for EImax have been given[28,31,35], containing the RMSA and either T1[35] (which can be assumed as proportional to PGV/PGA) or directly to PGV/PGA[28,31]. In this way, the ratio PGV/PGA can be assumed as a measure of the energy content of the earthquake[36]. The values of non-linear response parameters for some destructive earthquakes are given in Table 2, where Ehmaxis evaluated for = 4.

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS


collapse under cyclic loads, and introducing the damage index, D[68]. The assessment of intermediate levels of damage is an aspect which is more complex than the simple definition of collapse under cyclic actions and this is a topic where research is still ongoing. The modelling of the structural response under seismic actions of elements and of entire structures depends on structural type, materials and load history. The analysis of structural damage due to severe cyclic actions can be performed using a micromodel or a macromodel approach with different levels of refinement and consequently of performance. The macromodel approach simulates the response of structural elements using a lumped plasticity approach with hysteretic models or analyzes the entire structure as a SDOF system. It allows a simplified structural analysis whilst the main aspects of cyclic damage are considered. The definition of structural macromodels under cyclic loading can be developed by implementation of the following steps: G definition of a monotonic curve; the cyclic envelope can be coincident with the monotonic one or subject to degradation, G definition of the unloading rules; in the simplest case, the unloading branch can be linear, G definition of the re-loading rules; the re-loading branch may follow complex rules with slip, pinching, etc. On the basis of previous definitions, the following classes of hysteretic models can be distinguished[37,38]: Class I : non-evolutionary models, Class II : evolutionary models, Class III : degrading models. In order to clarify this classification, for the sake of simplicity, an elastoplastic force (F) displacement () relationship can be assumed as the monotonic envelope. The model is defined by the yield force, Fy, and the yield displacement, y. By definition, the state of the system is represented by the value of the displacement, , and of the velocity, d/dt, at time, t. The sign of the product d/dt indicates the loading or unloading condition.
Non-evolutionary models

Damage index and hysteretic behaviour


The damage analysis of seismic response requires the introduction of efficient structural models capable of describing actual behaviour. Conceptually, the modelling of the experimental behaviour of elements and structures under cyclic loading is based on two key points: G the definition of collapse criteria for random cyclic loads, and G the introduction of hysteretic rules in the constitutive relationships. These two points and the correlation between them will be discussed in the following. The definition of collapse criteria caused by seismic actions can be developed by identifying the structural parameters, di, which control collapse, determining the values of di at the damage threshold and at
Copyright 2000 John Wiley & Sons, Ltd

In the non-evolutionary models, the state of the system at time ti+1 only depends on the state at time, ti. Examples of non-evolutionary models are the elasticperfectly-plastic (EPP) and elastoplastic with hardening (EPH) or softening (EPS) models.
Evolutionary models

In the evolutionary models, the state of the system at time ti+1 depends on the state at various earlier times tj. Examples of evolutionary models are the originProg. Struct. Engng Mater. 2000; 2: 5059

DAMAGE INDICES AND DAMAGE MEASURES


(a) (b)

55

(c)

(d)

Fig. 2 Hysteretic behaviour: (a) elastoplastic model, (b) evolutionary model (CloughJohnston model), (c) degrading model (stiffness degradation), (d) degrading model (strength degradation)

oriented model, the Clough-Johnston model, the Slip model, the Q-model, etc.
Degrading models

In the degrading models, the state of the system at time ti+1 depends on the states at various earlier times tj and on the values at the time ti of appropriate damage indices. Different degrading models have been proposed for RC or steel structures[3742]. Examples of different hysteretic models are shown in Fig. 2. However, it is interesting to note that in some cases the damage index governing model degradation also governs collapse, but in other cases degradation and collapse depend on different damage indices.

Damage indices versus non-linear behaviour


The damage and collapse indices proposed in the scientific literature are numerous and various, and can be defined for each structural element or sub-elements (local indices) or related to the entire global structure (global indices). The local damage index is calibrated for each element (beams, columns, joints, shear walls, etc.) and for each loading condition (axial load, flexure, shear). The experimental identification is generally based on
Copyright 2000 John Wiley & Sons, Ltd

the statistical regression of experimental tests, while in some cases the experimental procedures are not standardized. Different papers have been published recently to detail the different proposed formulations[6,10], mainly regarding RC structures. The most commonly used local indices will be briefly described in the following sections in order to present some case studies. The structural quantities applied to the structures will be indicated in eqns (6) and (7) by the subscript dem, while the supplied quantities will be indicated by the subscript sup. The first applied structural damage parameter is the kinematic or cyclic ductility, , that can be defined in terms of rotation, curvature or displacement. The choice of the kinematic or cyclic ductility as a damage measure is equivalent to assuming that the collapse of the structural model is expected for maximum plastic displacement, independent of the number of plastic cycles and the amount of dissipated energy. Experimental tests have shown that collapse depends on the maximum displacement demand for well-detailed RC elements and structures without bond or shear failure and for steel structures without local buckling phenomena. However, it has been shown that this index can also be used for structures with cumulative deterioration such as in the case of impulse-type or short-duration earthquakes which
Prog. Struct. Engng Mater. 2000; 2: 5059

56
are characterized by one cycle with a large plastic displacement and others with a small amount of plastic work. The plastic dissipated energy, Eh, was first adopted as a damage measure considering cumulative energy absorption[34]. Different researchers have proposed variations of the index, including modification factors, to take into account the specific geometrical and loading conditions of RC[44,45] or steel elements. Experimental evaluation of the capacity of energy absorption is reported in ref.[46]. An improvement to this basic index has been proposed in ref.[47] introducing a more complex rule in the accumulation of energy. The energy-based indices are very attractive for their simplicity in the evaluation of dissipated energy in a structure as observed above, but the experimental assessment of the supplied energy dissipation capacity is very difficult. Another approach is based on the accumulation of damage due to cyclic loading that is usually modelled by introducing the low-cycle fatigue law. A general deformation cumulative index can be defined by introducing the well-known Coffin & Mason[48,49] and Miner[50,51] laws: (6) The value assumed by this damage index is defined in terms of constant b, which depends on the structural material and typology, and on the number of different plastic displacements, independent of the exact order of the displacements. Typical values of b, obtained by experimental data for steel structures[52] and RC structures[7,53], are 1.61.8; in a damage analysis, sometimes a conservative value of 1.5 is assumed[54]. Experimental values for different elements are recorded in ref.[55]. For b = 1 the index gives the same weight to each plastic displacement, independent of their individual values; hence, DF coincides with the energy-based index DE for the EPP model. For high values of b (ie > 5) the low-cycle fatigue index provides results similar to the ductility index[8]. The potential of the CoffinMason law to analyse low-cycle fatigue in steel components is shown in ref.[56], and in a RC section in ref.[53]. A generalized version of the cumulative index has been proposed in ref.[41] for RC structures considering separately positive and negative cycles and including weight factors in Miners rule. Different formulations of the cumulative index are proposed in refs[42,53] for concrete structures and in refs[57,58] for steel structures. The damage index based on low-cycle fatigue has the advantage of taking into account the distribution of the amplitude of plastic cycles and of having clear experimental tests to assess numerical parameters. Other indices are based on a combination of
Copyright 2000 John Wiley & Sons, Ltd

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS


ductility and dissipated energy demand. The most widely used index is the Park & Ang index[5961] which is defined as the linear combination of the maximum displacement and the dissipated energy, in the following way: (7) where, in RC structures, the parameter depends on the value of shear and axial forces in the section and on the total amount of longitudinal and confining reinforcement, by means of a regression curve obtained from more than 250 experimental results. The experimental values of reported in ref.[59] ranged between about -0.3 to +1.2, with a median of about 0.15[8]. There are interesting field applications for the index: the analysis of real buildings damaged by past earthquakes in Japan and in the USA is of remarkable interest. The analysis of the results has also showed that when D > 1 structural collapse occurred, while for D < 0.5 structural damage was not irreparable. In the case of 0.5 < D < 1.0, collapse did not occur, but the building could not be considered as repairable. On the other hand, the cases where damage is insignificant are defined by D < 0.2[61]. The index can take into account both maximum plastic displacement and plastic dissipated energy and is supported by a wide correlation with observed damage. However, the experimental determination of the parameter is difficult and the methodology is not well stated; another limitation is the linear combination of ductility and energy in a highly nonlinear problem. The index does not take into account the plastic cycles distribution, but considers only the global amount of dissipated energy. The Park & Ang index could assume values greater than 1. Chai et al[62] have recently proposed a rearrangement of the index setting the damage index to unity to signify structural failure. A more comprehensive index based on a combination of the Park & Ang index and the low-cycle fatigue index was proposed in ref.[38] for thin-walled steel bridge piers. Another combined index is that by Banon et al[54,63] with reference to the RC sections with combined axial load and bending. The analysis of seismic behaviour of the reference model EPP (non-evolutionary, non-degrading), extensively described in ref.[8], shows that (see also Fig. 3): G the scatter of results obtained when considering ductility or the plastic dissipated energy is very large, G the Park & Ang damage index with = 0.6 0.8 in practice provides the same result as the energy method; greater values of these parameters do not seem to have physical significance, G for very low values of (ie 0.050.10) energy dissipation in practice does not affect the final results compared with the ductility index,
Prog. Struct. Engng Mater. 2000; 2: 5059

DAMAGE INDICES AND DAMAGE MEASURES

57

Fig. 3 Inuence of damage parameters on seismic response (elastoplastic behaviour)

G a low-cycle fatigue damage index with b > 5, in practice, provides the same result as the ductility method; with higher values of this parameter it seems logical to use the simple ductility index method. G the Park & Ang and low-cycle fatigue damage indices, with suggested values of the parameters ( = 0.15 and b = 1.8), provide results which are, in practice, coincident. The analysis of seismic behaviour of the CloughJohnston, origin-oriented and Slip models, (nonevolutionary, degrading), compared with the response of the EPP model[37], shows (see Fig. 4): G the behaviour of structures governed by the ductility damage function seems to be only slightly influenced by the hysteretic models, G the behaviour of structures controlled by the energy damage function is more sensitive to modelling. The analysis of the evolutionarydegrading model proposed in ref.[37] shows that the effect of strength degradation and of re-loading stiffness degradation is

very important, and structures which show this kind of damage have to be designed using accelerations which are greater than those corresponding to nondegrading structures, while the effect of unloading stiffness degradation is less important. Therefore, it is worth pointing out that the structures characterized by strength degradation must be designed with particular attention; the structures with large slips due to bond and buckling of bars (RC structures) or with early local buckling (steel or composite structures) fall within this class. On the other hand, structures characterized only by unloading stiffness degradation can be simulated using non-degrading models like the CloughJohnston model, with small errors in the analysis.

Conclusions
Evaluation of damage measures and damage index is a very complex topic due to the complexity of the seismic input and the strong influence of structural response. In this review an attempt has been made at a systematic classification of recent papers on this subject.
(b)

(a)

Fig. 4 Inuence of hysteretic behaviour on seismic response: (a) functional cyclic ductility damage, (b) functional energy damage

Copyright 2000 John Wiley & Sons, Ltd

Prog. Struct. Engng Mater. 2000; 2: 5059

58 References and recommended reading


Papers of particular interest have been marked: * Special interest ** Exceptional interest
** [1] Fajfar P & Krawinkler H (eds). Nonlinear seismic analysis and design of reinforced concrete buildings. London: Elsevier Applied Science, 1992.

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS


[22] Vidic T, Fajfar P & Fischinger M. Consistent inelastic design spectra: strength and displacement. Earthquake Engineering and Structural Dynamics 1994: 23: 523532. [23] Housner G. Spectrum intensities of strong motion earthquakes. In: Proceedings Symposium on Earthquake and Blast Effects on Structures, Los Angeles, California, June 1952. 2036. [24] Kappos AJ. Sensitivity of calculated inelastic seismic response to input motion characteristics. In: Proceedings 4th US National Conference on Earthquake Engineering, Palm Springs, USA, 2024 May 1990. EERI. 1990. 2534. [25] Miranda E. Estimation of maximum interstory drift demands in displacement-based design. In: Fajfar P & Krawinkler H (eds) Seismic design methodologies for the next generation of codes. Rotterdam: Balkema. 1997. 253264. [26] Fajfar P, Vidic T & Fischinger M. A measure of earthquake motion capacity to damage medium period structures. Soil Dynamics and Earthquake Engineering 1990: 9(5): 236242. [27] Miranda E & Bertero VV. Evaluation of strength reduction factors for earthquake-resistant design. Earthquake Spectra 1994: 10(2): 357359. [28] Cosenza E & Manfredi G. The improvement of the seismic-resistant design for existing and new structures using damage criteria. In: Fajfar P & Krawinkler H (eds) Seismic design methodologies for the next generation of codes. Rotterdam: Balkema. 1997. 119130. [29] Zahrah T & Hall J. Earthquake energy absorption in SDOF structures. Journal of Structural Engineering (ASCE) 1984: 110(8): 17571772. [30] Rodriguez M. A measure of the capacity of earthquake ground motion to damage structures. Earthquake Engineering and Structural Dynamics 1994: 23: 627643. [31] Fajfar P & Vidic T. Consistent inelastic design spectra: hysteretic and input energy. Earthquake Engineering and Structural Dynamics 1994: 23: 523532. [32] Krawinkler H & Nassar AA. Seismic design based on ductility and cumulative damage demands and capacities. In: Fajfar P & Krawinkler H (eds) Nonlinear seismic analysis and design of reinforced concrete buildings. Oxford: Elsevier Applied Science. 1992. 2340. [33] Housner GW. Behaviour of structures during earthquake. Journal of the Engineering Mechanics Division, ASCE 1959: 85(4): 109129. [34] Akiyama H. Earthquake resistant limit-state design for buildings. Tokyo: University of Tokyo Press. 1985. [35] Kuwamura H & Galambos TV. Earthquake load for structural reliability. Journal of Structural Engineering (ASCE) 1989: 115(6): 14461462. [36] Tso WK, Zhu TJ & Heidebrecht AC. Seismic energy demands on reinforced concrete moment-resistant frames. Earthquake Engineering and Structural Dynamics 1993: 22: 533545. [37] Cosenza E & Manfredi G. Seismic analysis of degrading models by means of damage functions concept. In: Fajfar P & Krawinkler H (eds). Nonlinear seismic analysis and design of reinforced concrete buildings. Oxford: Elsevier Applied Science. 1992. 7794. [38] Kumar S & Usami T. An evolutionary-degrading hysteretic model for thin-walled structures. Engineering Structures 1996: 18(7): 504514. [39] Kunnath SK, Reinhorn AM & Park YJ. Analytical modeling of inelastic seismic response of R/C structures. Journal of Structural Engineering (ASCE) 1990: 116(4): 9961017. [40] Wen YK. Method for random vibration of hysteretic systems. Journal of Engineering Mechanics 1976: 102(2): 249263. [41] Chung YS, Meyer C & Shinozuka M. Seismic damage assessment of reinforced concrete members. Technical Report NCEER-87-0022. Buffalo, New York: National Center for Earthquake Engineering Research. 1987. [42] Wang ML & Shah SP. Reinforced concrete hysteresis model based on the damage concept. Earthquake Engineering and Structural Dynamics 1987: 15: 9931003. [43] Roufaiel MSL & Meyer C. Analytical modelling of hysteretic behaviour of R/C frames. Journal of Structural Engineering (ASCE) 1987: 113(3): 429444. [44] Gosain NK, Brown RH & Jirsa JO. Shear requirements for load reversals on RC members. Journal of Structural Engineering (ASCE) 1977: 103(7): 14611476. [45] Darwin D & Nmai CK. Energy dissipation in RC beams under cyclic load. Journal of Structural Engineering (ASCE) 1986: 112(8): 18291846. [46] Suzuki M, Akakura Y, Adachi H, & Ozaka Y. Evaluation of damage index for reinforced concrete structures. Concrete Library of JSCE 1995: 26: 117. [47] Kratzig WB, Meyer IF & Meskouris K. Damage evolution in reinforced concrete members under cyclic loading. In: Proceedings 5th International Conference on Structural Safety and Reliability (ICOSSAR 89), San Francisco, USA, 1989. Vol II. 795802. [48] Cofn LF. A study of the effect of cyclic thermal stresses in ductile metals. Transactions of the ASME 1954: 76: 931950. [49] Manson SS. Behaviour of materials under conditions of thermal stress. NASA TN 2933. Hanover, MD: NASA. 1954.

A collection of papers that represents the state of art on the nonlinear behaviour of structures under seismic loadings.
[2] Abrams DP & Calvi GM (eds). Proceedings USItalian Workshop on Seismic Evaluation and Retrot. Technical Report NCEER-97-0003. Buffalo, New York: National Center for Earthquake Engineering Research. 1997. [3] Bertero VV. State of art report on design criteria. In: Proceedings 11th World Conference on Earthquake Engineering, Acapulco, Mexico, 2328 June 1996. ** [4] Fajfar P & Krawinkler H (eds). Seismic design methodologies for the next generation of codes. Rotterdam: Balkema. 1997.

A collection of papers that represents the state of art in new trends in design methodologies related to the performance-based approach.
* [5] SEAOC Vision 2000 Committee. Performance based seismic design engineering. Sacramento, USA: Structural Engineers Association of California (SEAOC) Report. 1995.

The rst global proposal of a performance based design methodology.


** [6] Kappos AJ. Seismic damage index for RC buildings: evaluation of concepts and procedures. Progress in Structural Engineering and Materials 1997: 1(1): 7887.

This article gives a well-stated presentation of the different approach methodologies in the denition and evaluation of a damage index for RC structures.
[7] Powell GH & Allahabadi R. Seismic damage prediction by deterministic methods: concepts and procedures. Earthquake Engineering and Structural Dynamics 1988: 16: 719734. [8] Cosenza E, Manfredi G & Ramasco R. The use of damage functionals in earthquake-resistant design: a comparison among different procedures. Structural Dynamics and Earthquake Engineering 1993: 22: 855868. [9] McCabe SL & Hall WJ. Assessment of seismic structural damage. Journal of Structural Engineering (ASCE) 1989: 115(9): 21662183. [10] Williams MS & Sexsmith RG. Seismic damage indices for concrete structures: a state of the art review. Earthquake Spectra 1995: 11(2): 319349. [11] Fardis M. Damage measures and failure criteria for reinforced concrete members. In: Proceedings 10th European Conference on Earthquake Engineering, Vienna, 1995. Rotterdam: Balkema. 1995. 13771382. [12] Housner GW & Jennings PC. Earthquake design criteria. EERI Monograph Series. Berkeley, CA: Earthquake Engineering Research Institute. 1982. [13] Zhu TJ, Tso W K & Heidebrecht AC. Effect of peak ground a/v ratio on structural damage. Journal of Structural Engineering (ASCE) 1988: 114: 10191037. [14] Meskouris K, Kratzig WB & Hanskotter U. Seismic motion damage potential for R/C wall-stiffened buildings. In: Fajfar P & Krawinkler H (eds) Nonlinear seismic analysis and design of reinforced concrete buildings. Oxford: Elsevier Applied Science. 1992. 125136. [15] Arias A. A measure of earthquake intensity. In: Seismic design of nuclear power plants. Cambridge, MA: MIT Press. 1970. 438468. [16] Saragoni GR. Response spectra and earthquake destructiveness. In: Proceedings 4th US National Conference on Earthquake Engineering, Palm Springs, USA, 2024 May 1990. Berkeley, CA: Earthquake Engineering Research Institute (EERI). 1990. 3543. * [17] Uang CM & Bertero VV. Evaluation of seismic energy in structures. Earthquake Engineering and Structural Dynamics 1990: 19: 7790.

A basic paper on the energy approach in earthquake-resistant design.


[18] Trifunac MD & Brady AG. A study on the duration of strong earthquake ground motion. Bulletin of the Seismological Society of America 1975: 65(3): 581626. [19] Kawashima K & Aizawa K. Bracketed and normalized durations of earthquake ground acceleration. Earthquake Engineering and Structural Dynamics 1989: 18: 10411051. [20] Trifunac MD & Novikova EI. Duration of strong ground motion in terms of earthquake magnitude epicentral distance, site conditions and site geometry. Earthquake Engineering and Structural Dynamics 1994: 23: 10231043. [21] Newmark NM & Hall WJ. Earthquake spectra and design. EERI Monograph Series. 1982.

Copyright 2000 John Wiley & Sons, Ltd

Prog. Struct. Engng Mater. 2000; 2: 5059

DAMAGE INDICES AND DAMAGE MEASURES


[50] Miner MA. Cumulative damage in fatigue. Journal of Applied Mechanics 1945: 9. [51] Palmgren A. Die lebensdauer von kugellagern. Zeitschrift Vereines Deutscher Ingenieure 1924: 68. [52] Krawinkler H & Zohrei M. Cumulative damage in steel structures subjected to earthquake ground motion. Computers & Structures 1983: 16(14): 531541. [53] Stephens JE & Yao JTP. Damage assessment using response measurements. Journal of Structural Engineering (ASCE) 1987: 113(4): 787801. [54] Banon H, Biggs J & Irvine H. Seismic damage in reinforced concrete frames. Journal of Structural Engineering (ASCE) 1981: 107(9): 17131728. [55] Yamada M. Fracture ductility of structural elements and of structures. In: Proceedings 9th World Conference on Earthquake Engineering, Tokyo-Kyoto, Japan, 29 August 1988. Vol IV. Paper 6-3-3. [56] Krawinkler H. Performance assessment of steel components. Earthquake Spectra 1981: 3(1): 2741. [57] Ballio G & Castiglioni CA. An approach to the seismic design of steel structures based on cumulative damage criteria. Earthquake Engineering and Structural Dynamics 1994: 23: 969986.

59

[58] Calado L & Castiglioni CA. Steel beam-to-column connection under low cycle fatigue: experimental and numerical research. In: Proceedings 11th World Conference on Earthquake Engineering, Acapulco, Mexico, 2328 June 1996. [59] Park YJ. Seismic damage analysis and damage-limiting design of R/C structures. PhD Thesis, Department of Civil Engineering, University of Illinois, Urbana, IL. 1984. [60] Park YJ & Ang AH-S. Mechanistic seismic damage model for reinforced concrete. Journal of Structural Engineering (ASCE) 1985: 111(4): 722739. [61] Park YJ, Ang AH-S & Wen YK. Damage-limiting aseismic design of buildings. Earthquake Spectra 1987: 3(1): 126. [62] Chai YH, Romnstadt KM & Bird SM. Energy based linear damage model for high intensity seismic loading. Journal of Structural Engineering (ASCE) 1995: 121(5): 857864. [63] Banon H & Veneziano D. Seismic safety of reinforced concrete members and structures. Earthquake Engineering and Structural Dynamics 1982: 10: 179193.

Edoardo Cosenza PhD Professor, Dipartimento di Analisi e Progettazione Strutturale, Universit` a di Napoli Federico II, Via Claudio 21, 80125 Napoli, Italy E-mail: cosenza@unina.it Gaetano Manfredi PhD Associate Professor, Dipartimento di Analisi e Progettazione Strutturale, Universit` a di Napoli Federico II, Via Claudio 21, 80125 Napoli, Italy

Copyright 2000 John Wiley & Sons, Ltd

Prog. Struct. Engng Mater. 2000; 2: 5059

Das könnte Ihnen auch gefallen