Vous êtes sur la page 1sur 20

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS Earthquake Engng Struct. Dyn. 2004; 33:687706 (DOI: 10.1002/eqe.

369)

Behavior of moment-resisting frame structures subjected to near-fault ground motions


Babak Alavi1; ; and Helmut Krawinkler2
1 Exponent 2 Department

Failure Analysis Associates; 149 Commonwealth Drive; Menlo Park; CA 94025-1133; U.S.A. of Civil and Environmental Engineering; Stanford University; Stanford; CA 94305-4020; U.S.A.

SUMMARY Near-fault ground motions impose large demands on structures compared to ordinary ground motions. Recordings suggest that near-fault ground motions with forward directivity are characterized by a large pulse, which is mostly orientated perpendicular to the fault. This study is intended to provide quantitative knowledge on important response characteristics of elastic and inelastic frame structures subjected to near-fault ground motions. Generic frame models are used to represent MDOF structures. Near-fault ground motions are represented by equivalent pulses, which have a comparable e ect on structural response, but whose characteristics are dened by a small number of parameters. The results demonstrate that structures with a period longer than the pulse period respond very di erently from structures with a shorter period. For the former, early yielding occurs in higher stories but the high ductility demands migrate to the bottom stories as the ground motion becomes more severe. For the latter, the maximum demand always occurs in the bottom stories. Preliminary regression equations are proposed that relate the parameters of the equivalent pulse to magnitude and distance. The equivalent pulse concept is used to estimate the base shear strength required to limit story ductility demands to specic target values. Copyright ? 2004 John Wiley & Sons, Ltd.
KEY WORDS:

near-fault; near-eld; near-source, pulse; frame structures; seismic demands

INTRODUCTION Ground motion recordings have provided increasing evidence that ground shaking near a fault rupture may be characterized by a large, long-period pulse, capable of causing severe structural damage. This holds true particularly in the forward direction, where the fault
Correspondence E-mail:

to: Babak Alavi, Exponent Failure Analysis Associates, 149 Commonwealth Drive, Menlo Park, CA 94025-1133, U.S.A. balavi@exponent.com

Contract= grant sponsor: CUREE= Kajima Research Program Contract= grant sponsor: U.S. National Science Foundation, USJapan Cooperative Research Program in Urban Hazard Mitigation; contract= grant number: CMS-9812478 Contract= grant sponsor: California Department of Conservation, SMIP 1997 Data Interpretation Project; contract= grant number: 1097-601

Copyright ? 2004 John Wiley & Sons, Ltd.

Received 11 February 2003 Revised 11 September 2003 Accepted 3 October 2003

688

B. ALAVI AND H. KRAWINKLER

rupture propagates towards the site at a speed close to the shear wave velocity. As a result, most of the seismic energy from the rupture arrives within a short time at the beginning of the record [1]. The radiation pattern of shear dislocation around the fault causes the fault-normal component to be typically more severe than the fault-parallel component [2]. This phenomenon a ects the response attributes of structures located in the near-fault region, which is assumed to extend 10 to 15 km from the seismic source [3], and therefore requires consideration in the design process. Recent structural design codes, e.g. the 1997 Uniform Building Code [4], have recognized near-fault e ects by introducing source type and distance dependent near-fault factors to the customary design spectrum. However, these factors are believed to be inadequate to provide consistent protection because they pay little attention to the physical response characteristics of near-fault ground motions. Moreover, emerging concepts of performance-based seismic design require a quantitative understanding of response covering the range from nearly elastic behavior to highly inelastic behavior. Much work is needed to identify and quantify the sitedependent characteristics of near-fault ground motions and to address issues concerning the response of di erent types of structures to these ground motions. This paper summarizes the results of a study that aims to provide quantitative knowledge on the response of frame structures in the near-fault region of active faults. The objective is to identify salient response characteristics, to describe near-fault ground motions by simple equivalent pulses, and to use the pulse response characteristics in order to represent behavior attributes of structures subjected to near-fault ground motions. It is recognized that the nearfault problem is very complex, and that much more research and data are needed before a comprehensive understanding of all important aspects of the problem will be accomplished. The work summarized here attempts to address the most important issues concerning nearfault ground motions and their response attributes in order to form a foundation on which to base future research and development of design guidelines.

GROUND MOTIONS USED IN THIS STUDY A set of 15 near-fault ground motion records with forward directivity is used to evaluate elastic and inelastic demands of SDOF (single degree of freedom) and MDOF (multi degree of freedom) structures. These ground motions, which were assembled for the SAC Steel Project [5] and the CDMG Strong Motion Instrumentation Program [2], are either recorded on sti soil or have been modied to NEHRP [6] Soil Type D conditions. Table I lists the designation and basic properties of the recorded motions. These motions cover a moment magnitude range from 6.2 to 7.3 and a rupture distance (closest distance from site to fault rupture plane) range from 0.0 to 8:9 km. To augment the relatively small record set, a set of simulated ground motions is also used in the investigation, which covers systematic ranges of moment magnitude (6.5, 7.0 and 7.5) and rupture distance (3, 5 and 10 km) for two stations (f6 and f8) in the forward direction of a strike-slip fault. These records were generated for a project sponsored by the CDMG Strong Motion Instrumentation Program [2]. Records with backward directivity are typically less severe and do not exhibit pulse-type characteristics [1]; therefore, this study only focuses on ground motions with forward directivity. Figure 1 illustrates ground acceleration, velocity and displacement time history traces for the fault-normal component of a typical near-fault ground motion with forward directivity
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

689

Table I. Designation and basic properties of recorded ground motions used in this study.
Designation LP89lgpc LP89lex EZ92erzi LN92lucr NR94rrs NR94sylm KB95kobj KB95tato IV79ar06 IV79melo KB95kpi1 MH84andd MH84cyld NR94newh NR94spva Earthquake Loma Prieta, 1989 Loma Prieta, 1989 Erzincan, 1992 Landers, 1992 Northridge, 1994 Northridge, 1994 Kobe, 1995 Kobe, 1995 Imperial Valley, 1979 Imperial Valley, 1979 Kobe, 1995 Morgan Hill, 1984 Morgan Hill, 1984 Northridge, 1994 Northridge, 1994 Station Los Gatos Lexington Erzincan Lucerne Rinaldi Olive View JMA Takatori Array 6 Meloland Port Island Anderson D Coyote L D Newhall Sepulveda Mw 7.0 7.0 6.7 7.3 6.7 6.7 6.9 6.9 6.5 6.5 6.9 6.2 6.2 6.7 6.7 R (km) 3.5 6.3 2.0 1.1 7.5 6.4 0.6 1.5 1.2 0.0 3.7 4.5 0.1 7.1 8.9

NR94rrs, Fault-Normal 1000


max = 873

ag (cm / sec2)

500

-500

-1000

200
max = 174

vg (cm / sec)

100

-100

-200

40
max = 38.3

20

ug (cm)

-20

-40 0 2 4 6 8 10 12 14

Time (sec)

Figure 1. Ground acceleration, velocity and displacement time histories for record NR94rrs.
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

690

B. ALAVI AND H. KRAWINKLER


NR94rrs, = 2% 3
Fault-Normal

2.5 2

Fault-Parallel 0.707(FN+FP) 0.707(FN-FP)

Sa (g)

1.5 1 0.5 0

400
Fault-Normal Fault-Parallel

300

Sv (cm / sec)

0.707(FN+FP) 0.707(FN-FP)

200

100

100 80
Fault-Normal Fault-Parallel 0.707(FN+FP) 0.707(FN-FP)

Sd (cm)

60 40 20 0 0 0.5 1 1.5 2 2.5 3 3.5 4

T (sec)

Figure 2. Acceleration (elastic strength demand), velocity and displacement spectra for record NR94rrs.

(NR94rrs = Northridge 94 Rinaldi Receiving Station). As indicated particularly by the velocity and displacement traces, the record contains a large pulse within the time range from about 2 to 3 sec. Figure 2 presents acceleration (elastic strength demand), velocity and displacement spectra for the fault-normal, fault-parallel and two 45-degree rotated components of the same ground motion. The gure indicates that the fault-normal component is much more severe than the fault-parallel component. When these two components are rotated by 45 degrees, the di erence in the spectra becomes smaller, but one of the two rotated components still will impose demands close to (and sometimes even higher than) those associated with the fault-normal component. A more comprehensive study of the orientation e ect conrmed this observation [7]. Thus, when a structure composed of frames in two perpendicular directions is subjected to a near-fault ground motion, frames in one of the two directions will be subjected to excitations almost as severe as the fault-normal component. For this reason this study focuses on the fault-normal component of near-fault ground motions. To put the severity of near-fault ground motions in perspective with the severity of ground motions represented by current codes, a reference set of 15 ordinary records (recorded outside the near-fault region) is used for comparison purposes. These records, which were used in past studies [8], are scaled such that the spectrum of each individual record matches
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

691

Elastic SDOF Strength Demand Spectra


2 1.6 1.2 0.8 0.4 0 0 0.5 1 1.5 2 2.5 3 3.5 4
15-D* Records and UBC 97 Soil Type SD, = 5%

UBC 97 15-D* (mean)

Sa (g)

T (sec)

Figure 3. Mean acceleration (elastic strength demand) spectrum for reference record set, 15-D*.
Elastic SDOF Velocity Demands
800
15-D* vs. Recorded Near-Field, = 2%
LP89lgpc EZ92erzi NR94rrs KB95kobj 15-D* (mean) LP89lex LN92lucr NR94sylm KB95tato

600

Sv (cm / sec)

400

200

0 0 0.5 1 1.5 2 2.5 3 3.5 4

T (sec)

Figure 4. Velocity response spectra for near-fault and reference ground motions.

the UBC 97 soil type SD spectrum with a minimum error, using discrete periods in the range of 0.6 to 4:0 sec (constant velocity range). The mean acceleration spectrum of the 15 scaled records, referred to as 15-D* (mean), is shown in Figure 3 together with the UBC 97 spectrum for Seismic Zone 4 (without a near-fault factor). Thus, on average the 15-D* time histories reasonably represent the UBC design level. Figure 4 illustrates the mean velocity spectrum of the 15-D* records superimposed on the velocity spectra of several of the recorded near-fault time histories. This gure is presented for two reasons: rst, to illustrate the great variability in near-fault response spectra and second, to put the severity of near-fault ground motions in perspective with present design ground motions. The gure demonstrates that near-fault records can impose very large demands that should be taken into account in the design process. An important observation from the near-fault spectra (Figures 2 and 4) is the existence of a predominant peak in the velocity spectrum of most of the near-fault recordsalthough some of the records have more than one clear velocity peak. The predominant peak of the velocity spectrum is used later to estimate the period of the pulse contained in the near-fault record.
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

692

B. ALAVI AND H. KRAWINKLER

MDOF SYSTEMS USED IN THIS STUDY In order to quantify seismic demands, MDOF structures are represented by a generic 20-story, single-bay frame whose fundamental elastic period and base shear strength are varied. The base shear strength is dened by the coe cient = Vy =W , where Vy is the base shear strength and W is the weight of the structure. The story sti nesses and shear strengths are determined from a lateral load pattern that corresponds to story shear forces resulting from the SRSS combination of modal responses for a constant velocity response spectrum, e.g., the UBC design spectrum for T Ts [4]. This load pattern, which will be referred to as the SRSS pattern, is in line with current seismic design load patterns. The variation of the moments of inertia of the frame members over the height is tuned such that a linear static analysis of the frame subjected to the SRSS load pattern results in a straight-line de ected shape. The bending strengths of the elements are tuned such that under the SRSS pattern all plastic hinges, which are allowed only at ends of the beams and at the base of the columns, form simultaneously. Thus, the story shear strength pattern follows the same distribution as the design SRSS pattern, which implies a constant overstrength factor for each story. A bilinear non-degrading hysteresis model with a 3% strain-hardening ratio is used at all plastic hinge locations. In the time history analyses, Rayleigh damping is used to obtain a damping ratio of 2% at the rst mode period T and at 0:1T . The low damping is used because the emphasis is on the behavior of steel frame structures, and because the maximum response of structures subjected to pulse-type excitations is relatively insensitive to the damping ratio [9]. This paper does not address P-delta e ects. However, it must be emphasized that other results of this study have demonstrated that P-delta may have a signicant e ect on structural response to near-fault ground motions and generally deserve serious consideration. The secondorder e ects may be important even in cases in which current design provisions do not require any measures to account for P-delta e ects [7].

RESPONSE TO NEAR-FAULT GROUND MOTIONS The story ductility ratio, dened as the maximum story drift normalized by the story yield drift, i.e. i = max; i = y; i , is used to quantify the response of MDOF structures to near-fault ground motions. Figure 5 illustrates distributions of story ductility demands over the height of the MDOF structure previously introduced subjected to the near-fault records whose velocity response spectra are shown in Figure 4. The demands are computed for MDOF systems with a fundamental period T = 2:0 sec. and base shear strength coe cients of = 0:4 and = 0:15, which correspond to a relatively strong and a relatively weak structure, respectively. The relatively long period of 2:0 sec. is chosen because, as will be shown later, the period of the pulse contained in most of the near-fault records is shorter than 2:0 sec. For comparison purposes, the mean story ductility demands obtained from the reference record set 15-D* are superimposed. An important observation is that for most of the near-fault records, the maximum story ductility demand occurs in the upper portion of the structure when the structure is strong (large ). However, a migration of ductility demands toward the base takes place when the structure becomes weaker (small ). The non-uniformity of the story ductility demands
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

693

Story Ductility Demands


1 0.8
15-D* vs. Recorded Near-Field, T = 2.0 sec, = 0.40
LP89lgpc LP89lex EZ92erzi LN92lucr NR94rrs NR94sylm KB95kobj KB95tato 15-D* (mean)

Story Ductility Demands


1 0.8
15-D* vs. Recorded Near-Field, T = 2.0 sec, = 0.15
LP89lgpc LP89lex EZ92erzi LN92lucr NR94rrs NR94sylm KB95kobj KB95tato 15-D* (mean)

Relative Height

0.6 0.4 0.2 0 0 1 2 3 4

Relative Height
5

0.6 0.4 0.2 0 0 2 4 6 8

10

12

(a)

Story Ductility Ratio, i = max,i / y,i

(b)

Story Ductility Ratio, i = max,i / y,i

Figure 5. Story ductility demands for several near-fault and reference ground motions, T = 2:0 sec: (a) = 0:40; and (b) = 0:15.

Story Ductility Demands


1 0.8
NR94rrs, T = 0.5 sec, without P- = 2.00 = 1.50 = 1.00 = 0.80 = 0.60 = 0.50 = 0.40 = 0.30 = 0.25

Story Ductility Demands


1 0.8
NR94rrs, T = 2.0 sec, without P- = 0.80 = 0.50 = 0.40 = 0.25 = 0.20 = 0.15 = 0.10 = 0.07 = 0.05

Relative Height

0.6 0.4 0.2 0 0 4 8 12 16

Relative Height
20

0.6 0.4 0.2 0 0 4 8 12 16

20

(a)

Story Ductility Ratio, i = max,i / y,i

(b)

Story Ductility Ratio, i = max,i / y,i

Figure 6. Dependence of distribution of story ductility demands on base shear strength for record NR94rrs: (a) T = 0:5 sec; and (b) T = 2:0 sec.

observed for individual near-fault records holds true also in the mean, whereas the same SRSS-based story shear strengths on average result in a relatively uniform ductility distribution for ordinary ground motions (15-D*). Figure 5 also indicates the severity of near-fault ground motions in comparison with mean demands for the ordinary ground motions scaled to the UBC spectrum. The strength and period dependence of story ductility distributions is illustrated in Figure 6 for the record NR94rrs. The distributions are shown for relatively short (T = 0:5 sec:) and relatively long (T = 2:0 sec:) period structures whose base shear strength coe cient, , is varied progressively. As with the previous observation from Figure 5, for the structures with T = 2:0 sec. (a period that clearly exceeds the e ective pulse period of 1:0 sec. for NR94rrs), maximum ductility demands occur in the upper stories when the structure is strong. With a reduction in the base shear strength, the ductility demands in the upper portion stabilize and grow no more. Further strength reductions result in a migration of the maximum demand
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

694

B. ALAVI AND H. KRAWINKLER

Normalized Elastic Story Shear Demands


NR94rrs, without P-, Time History

1 0.8 Relative Height 0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 Vi,max / Vbase,max 1 1.2

Design T = 0.5 sec T = 1.0 sec T = 1.5 sec T = 2.0 sec T = 2.5 sec T = 3.0 sec T = 4.0 sec

Figure 7. Normalized elastic story shear demands for record NR94rrs.


Maximum MDOF Story Ductility Demands
2
NR94rrs, SRSS Pattern, without P-
T = 0.38 sec T = 0.50 sec

1.5

T = 0.75 sec T = 1.00 sec

= Vy / W

T = 1.50 sec

T = 2.00 sec T = 3.00 sec

0.5

0 0 4 8 12 16 20

Maximum Story Ductility, max

Figure 8. Base shear strength vs. maximum story ductility demands for record NR94rrs.

toward the base. This phenomenon is not observed for the structures with T = 0:5 sec:, a period that is shorter than the e ective pulse period. For these structures the maximum ductility demands occur close to the base regardless of the base shear strength. The reason for the early inelastic behavior in the top portion of long-period, strong structures lies in the distribution of elastic story shear demands. Figure 7 illustrates such distributions over the height of structures with various periods subjected to the record NR94rrs. Superimposed is also the SRSS story shear strength distribution used to design the MDOF systems (denoted as Design). All story shears are normalized by their corresponding base shear values. While the shear distributions for short-period systems (T 61:0 sec:) are relatively smooth and resemble the design distribution, their long-period counterparts exhibit the effect of a wave traveling up the structure, causing distributions that deviate considerably from the design shear distribution [7]. The elastic story shear demands are particularly large in the upper portion of long-period structures (T 1:0 sec:), where the shear demands reach the corresponding story shear capacity rst, resulting in early yielding in the upper stories. A comprehensive assessment of maximum story ductility demands (maximum over all stories) of the generic MDOF frames subjected to the near-fault record NR94rrs can be obtained from the max curves presented for various structure periods in Figure 8. It is noted that the
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

695

1.5 1

1.5 1

ag / ag,max

ag / ag,max

0.5

0.5 0

-0.5 -1 -1.5 1.5 1

-0.5 -1 -1.5 1.5

vg,max = ag,max.Tp / 4

vg,max = ag,max.Tp / 4

vg / vg,max

vg / vg,max

0.5 0 -0.5 -1 -1.5

0.5 0 -0.5 -1 -1.5

1.5 1

1.5

ug,max = ag,max.Tp2 / 16

ug,max = ag,max.Tp2 / 32

ug / ug,max

0 -0.5 -1 -1.5 0 0.5 1 1.5 2 2.5 3 3.5

ug / ug,max

0.5

0.5 0 -0.5 -1 -1.5 0 0.5 1 1.5 2 2.5 3 3.5

(a)

t / Tp

(b)

t / Tp

Figure 9. Acceleration, velocity and displacement time histories of basic pulses: (a) pulse P2; and (b) pulse P3.

max curves for systems with T = 1:5 and 2:0 sec. include a range with a very steep slope at 4, which corresponds to the migration of ductility demands from the upper stories to the = bottom story (see Figure 6(b)). The steep slope implies that for a signicant range of base shear strength (represented by ) the maximum story ductility demand remains about the same. RESPONSE TO PULSE INPUT As pointed out in the introduction, ground shaking in the forward directivity region of a fault rupture is characterized by a large pulse early in the time history. If simple pulse models can be introduced to represent near-fault ground motions with reasonable accuracy, the process of design and response evaluation will be greatly facilitated because of common patterns for structures whose fundamental period T has a specic relation to the pulse period Tp . This study focuses on three basic pulses for response evaluations: half pulse (P1), full pulse (P2) and multiple pulse (P3). Ground acceleration, velocity and displacement time histories for pulses P2 and P3 are shown in Figure 9. Pulse P1, whose time histories are not shown, consists of only the rst half of pulse P2. These pulses are fully dened by a pulse shape and two parameters: the pulse period Tp (duration of a full velocity cycle) and an intensity measure, which can be either the maximum pulse acceleration, ag; max , or the
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

696
8

B. ALAVI AND H. KRAWINKLER

6
g,max

P1 P2 P3

S /a

0 5 4
g,max

P1 P2 P3

3 2 1 0 6 5 4
P1 P2 P3

g,max d

S /V S /u

3 2 1 0 0 0.5 1 1.5 2 2.5 3

T / Tp

Figure 10. Acceleration (elastic strength demand), velocity and displacement spectra for basic pulses.

maximum pulse velocity, vg; max (= ag; max Tp = 4). The authors have shown that these square-wave acceleration pulses are adequately capable of representing other pulse shapes such as triangular acceleration pulses [7]. Other researchers have also investigated various pulse shapes such as sine-wave acceleration pulses to represent both fault-normal and fault-parallel components of ground motions with forward directivity [1012]. The great advantage of the square pulse is the unambiguous denition of maximum acceleration and the simple relationship between maximum acceleration and maximum velocity. The elastic response spectra for the three basic pulses are presented in Figure 10. The period axis is normalized by the pulse period Tp , and the spectral ordinates are normalized by their corresponding peak time history value. The velocity spectra for P2 and P3 show a peak at T=Tp = 1:0, indicating that the maximum spectral velocity for these pulses occurs at a period that matches well with the pulse period. This property can be taken advantage of to estimate the equivalent pulse period for near-fault records. The spectra also show larger spectral peaks for pulse P3 than for the other two pulses, particularly in the displacement domain. But in this domain it must be considered that for a given pulse intensity (e.g. ag; max ), the peak ground displacement for P3, used to normalize the displacement spectrum, is only half of that for P2 (see Figure 9). Nevertheless, it is evident that the multiple pulse P3 is more damaging than a single pulse.
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

697

Story Ductility Demands 1 0.8


Pulse P2, SRSS Pattern, T / Tp = 0.5, without P- = 2.50 = 2.00 = 1.50 = 1.25 = 1.00 = 0.75 = 0.60 = 0.50 = 0.40

Story Ductility Demands 1 0.8


Pulse P2, SRSS Pattern, T / Tp = 2.0, without P- = 1.25 = 1.00 = 0.75 = 0.50 = 0.25 = 0.15 = 0.10 = 0.07 = 0.05

Relative Height

Relative Height

0.6 0.4 0.2 0 0 4 8 12 16

0.6 0.4 0.2 0

20

12

16

20

(a)

Story Ductility Ratio, i = max,i / y,i

(b)

Story Ductility Ratio, i = max,i / y,i

Figure 11. Dependence of distribution of story ductility demands on base shear strength for pulse P2: (a) T=Tp = 0:5; and (b) T=Tp = 2:0.

The generic MDOF frame previously introduced is used to study the distributions of story ductility demands when the structure is subjected to the basic pulses. The base shear strength in this case is dened by the coe cient = Vy = (m:ag; max ), which expresses the base shear strength relative to the pulse intensity. Figure 11 illustrates story ductility distributions for pulse P2 and period ratios T=Tp = 0:5 and 2.0, representing structures with fundamental periods clearly shorter and longer than the pulse period. The distributions can be directly compared to those presented in Figure 6 for the near-fault record NR94rrs, whose Tp is 1:0 sec. This comparison reveals notable similarities between the response to the record NR94rrs and the response to pulse P2. These consistent similarities between MDOF response properties of near-fault ground motions with forward directivity and those of the basic pulses support the argument that near-fault records can indeed be represented by equivalent pulses. A comprehensive assessment of maximum story ductility demands (maximum over all stories) can be obtained from the max diagrams shown for pulse P2 and various T=Tp ratios in Figure 12. For T=Tp = 1:5 to 3.0, there is a range in which (base shear strength) can be reduced in half without leading to an increase in the maximum ductility demands. This close-to-vertical range, which is located in the ductility range of 3 to 4, corresponds to the migration of the maximum ductility demand from upper stories to the bottom story. Once the maximum ductility demand has reached the bottom story, it increases rather rapidly with a further reduction in strength. This pattern is in agreement with the observations made for near-fault ground motions (see Figure 8). In design it is often more useful to rearrange the information presented in Figure 12 in order to evaluate the base shear strength, , required to limit the maximum story ductility demand, max , to specic target values. Vertical cuts through the max diagram (supplemented by a linear interpolation scheme) provide values for the MDOF strength demands shown in Figure 13. This gure represents MDOF base shear strength demand spectra for pulse P2 and target maximum story ductility ratios of max = 1, 2, 3, 4, 6 and 8. Inherent in these spectra is the assumption that the story shear strength distribution over the height follows the SRSS story shear distribution. Provided that the basic pulses introduced in this study can represent near-fault ground motions, these MDOF spectra can be used for design against near-fault e ects, as will be shown later.
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

698

B. ALAVI AND H. KRAWINKLER

Maximum MDOF Story Ductility Demands 5 4


Pulse P2, SRSS Pattern, without P- T/Tp = 0.375 T/Tp = 0.50 T/Tp = 0.75 T/Tp = 1.00 T/Tp = 1.50 T/Tp = 2.00 T/Tp = 3.00

= Vy / (m.ag,max)

3 2 1 0 0 4

Maximum Story Ductility, max

12

16

20

Figure 12. Base shear strength vs. maximum story ductility demands for various T=Tp values, pulse P2.
MDOF Strength Demands for Constant Ductility
Pulse P2, SRSS Pattern, without P-

5 4
=1 =2 =3 =4 =6 =8

= Vy / (m.ag,max)

3 2 1 0 0 0.5 1 1.5 2 2.5

T / Tp

Figure 13. MDOF base shear strength demand spectra for target maximum story ductility ratios.

EQUIVALENT PULSES FOR NEAR-FAULT GROUND MOTIONS Since near-fault ground motions come in great variations (as evidenced by the responses shown in Figures 4 and 5), any attempt to evaluate or predict structural response will be complex unless near-fault ground motions can be represented by simplied motions that reasonably replicate important near-fault response characteristics. An inspection of near-fault time history traces, as well as the study of similarities between the response of structures subjected to near-fault records and simple pulses, provides evidence thatwithin limitationsnear-fault records with forward directivity may be represented by equivalent pulses of the type introduced previously. It is, however, unreasonable to expect that this equivalence can be established accurately for the full period range of interest. In the very short period range, a near-fault record is likely to be contaminated by high-frequency components that have little to do with characteristics of the long-period high-energy pulse generated by the propagation of fault rupture. In the very long period range, it is also likely that other phenomena (e.g., basin e ects and instrument errors) contaminate the record. In this study it is postulated that an equivalence between a near-fault ground motion and a pulse can be reasonably established in a range of T=Tp from
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

699

0.375 to 3.0. This range is based on a comparison between spectral shapes for near-fault records and the basic pulses, giving consideration to the range in which Tp varies for typical near-fault ground motions. It should be emphasized that the equivalent pulse is by no means a precise representation of near-fault ground motions. Rigorous procedures for the identication of equivalent pulse parameters have been established, but in some cases judgmental decisions need to be employed in order to arrive at nal values for these parameters [7]. In order to establish an equivalent pulse for a record, three parameters need to be evaluated: the pulse type (P1, P2 or P3), pulse period Tp , and pulse intensity ag; max (see Figure 9). An inspection of time history traces and a comparison between ground motion and pulse spectral shapes are employed to decide on the pulse type. The pulse period Tp is identied as the period corresponding to a global peak in the velocity response spectrum. The basis for this approach is that both P2 (full cycle) and P3 (multiple cycle), which represent the near-fault ground motions used in this study, show a hump at T=Tp = 1 (see Figure 10). While in many cases a narrow range for Tp can be identied, there are near-fault records whose velocity spectra include a wide peak or multiple peaks (see Figure 4). In such cases judgment has to be employed to decide on a nal value. The authors have performed a sensitivity study to evaluate these judgmental decisions and have come to the conclusion that, within limitations, the structural response is not very sensitive to the choice of Tp , provided that the intensity matching follows the procedure outlined in the next paragraph [7]. To estimate the intensity of the equivalent pulse, a rigorous process is employed whose objective is to minimize the di erences between the maximum story ductility demands obtained from the near-fault record and the equivalent pulse. In the following discussion the maximum acceleration, ag; max , and velocity, vg; max , of the equivalent pulse are referred to as the e ective acceleration, ae , and velocity, ve . In summary, the procedure includes the following steps: 1. Compute max curves (see Figure 12) for the appropriate pulse type and T=Tp = 0:375, 0.5, 0.75, 1.0, 1.5, 2.0 and 3.0. 2. Compute max curves for the near-fault record (see Figure 8) and T=Tp = 0:375, 0.5, 0.75, 1.0, 1.5, 2.0 and 3.0. 3. For each T=Tp value, convert the pulse max curve into a max curve [ = Vy = (m:g) = (ae =g)] and nd best-t values for ae by minimizing the relative di erences between the pulse and ground motion max curves for 16 max 610 (from nearly elastic to highly inelastic behavior). The di erences between the two curves are minimized using the least-squares method. 4. Obtain nal values for ae by averaging the best-t values for the seven period ratios. Tables II and III summarize the results of this procedure for the recorded and simulated near-fault ground motions, respectively. The Landers record (LN92lucr) is omitted from Table II because its pulse period is longer than 4 sec. (see Figure 4), and the matching procedure for this record would involve computing demands for structures with unreasonably long periods (e.g., T = 3Tp 12:0 sec:). The tables list the pulse type, pulse period, and maximum acceleration and velocity of the equivalent pulse. The peak velocity of the pulse is computed from the equation ve = ae Tp = 4. Comparing the so-computed ve with the recorded peak ground velocity (PGV) listed in the last column of Tables II and III indicates that in most of the cases the peak velocity of the equivalent pulse is very close to the PGV of the near-fault record. Only for the MH84cyld record does the di erence exceed 20%. Thus, it appears to be feasible to use the PGV of the
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

700

B. ALAVI AND H. KRAWINKLER

Table II. Equivalent pulses for recorded ground motions.


Designation LP89lgpc LP89lex EZ92erzi NR94rrs NR94sylm KB95kobj KB95tato IV79ar06 IV79melo KB95kpi1 MH84andd MH84cyld NR94newh NR94spva Mw 7.0 7.0 6.7 6.7 6.7 6.9 6.9 6.5 6.5 6.9 6.2 6.2 6.7 6.7 R (km) 3.5 6.3 2.0 7.5 6.4 0.6 1.5 1.2 0.0 3.7 4.5 0.1 7.1 8.9 Pulse type P3 P2 P2 P2 P3 P3 P3 P2 P2 P2 P2 P3 P2 P3 Tp (sec) 3.0 1.0 2.3 1.0 2.4 0.9 2.0 3.4 2.8 1.8 0.8 0.8 1.3 2.7 ae (g) 0.23 0.70 0.17 0.72 0.18 0.86 0.40 0.13 0.15 0.25 0.16 0.47 0.37 0.09 ve (cm/sec) 169 172 96 177 106 190 196 108 103 110 31 92 118 60 PGV (cm/sec) 173 179 119 174 122 160 174 110 117 100 27 65 119 63

Table III. Equivalent pulses for simulated ground motions.


Station f6 f8 f6 f8 f6 f8 f6 f8 f6 f8 f6 f8 f6 f8 f6 f8 f6 f8 Mw 6.5 6.5 6.5 6.5 6.5 6.5 7.0 7.0 7.0 7.0 7.0 7.0 7.5 7.5 7.5 7.5 7.5 7.5 R (km) 3 3 5 5 10 10 3 3 5 5 10 10 3 3 5 5 10 10 Pulse type P2 P2 P2 P2 P2 P2 P2 P2 P2 P2 P2 P2 P3 P3 P3 P3 P3 P3 Tp (sec) 1.7 1.2 2.0 2.1 2.6 3.0 3.2 3.4 3.5 3.6 5.0 3.3 3.2 3.2 3.2 3.2 3.2 3.2 ae (g) 0.17 0.41 0.10 0.19 0.03 0.05 0.13 0.21 0.11 0.16 0.04 0.10 0.21 0.27 0.19 0.24 0.14 0.18 ve (cm/sec) 71 121 49 98 19 37 102 175 94 141 49 81 165 212 149 188 110 141 PGV (cm/sec) 68 116 46 70 25 32 98 146 89 124 51 76 148 210 138 201 84 131

near-fault record to estimate the pulse intensity parameter (i.e., ae = 4PGV=Tp ) rather than following the elaborate procedure outlined previously. Examples of story ductility distributions over the height for the NR94rrs record and its equivalent pulse are shown in Figure 14 for cases of strong and weak structures. Although some di erences exist, the equivalent pulse appears to reasonably capture the important response characteristics of the near-fault record, particularly the migration of ductility demands from the top to the bottom portion of the long-period structure (T=Tp = 2:0) if the strength of the structure is decreased.
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

701

Story Ductility Demands


1 0.8
Low Ductility, T / Tp = 2.0, = 0.4, without P- NR94rrs

Story Ductility Demands


1 0.8
High Ductility, T / Tp = 2.0, = 0.1, without P- NR94rrs Pulse P2

Relative Height

0.6 0.4 0.2 0 0 1 2 3 4

Relative Height

Pulse P2

0.6 0.4 0.2 0 0 2 4 6 8 10 12

(a)

Story Ductility Ratio, i = dyn,i / y,i

(b)

Story Ductility Ratio, i = dyn,i / y,i

Figure 14. Comparison of story ductility demands obtained from record NR94rrs and its equivalent pulse, T=Tp = 2:0: (a) high strength, low ductility; and (b) low strength, high ductility.

MAGNITUDE AND DISTANCE DEPENDENCE OF EQUIVALENT PULSE PROPERTIES As Table II and, more systematically, Table III indicate, the parameters of the equivalent pulse for a near-fault record, i.e. Tp and ve (or ae ), appear to be related to the magnitude of the event (Mw ) and the closest distance from the rupture plane (R). Somerville showed that Tp is mostly a ected by slip rise time, dened as the duration of slip at a given point on the fault, which in turn is related to Mw [2]. He proposed a preliminary model that expresses log10 Tp as a linear function of Mw , independent of distance. Employing his functional form for Tp and performing a linear regression analysis on the equivalent pulse periods obtained for the combination of recorded and simulated ground motions (Tables II and III), the following regression equation is obtained: log10 Tp = 1:76 + 0:31Mw (1)

Figure 15 illustrates this equation together with the data points (solid circles) to which the line is tted. Some of the circles represent multiple data points with identical Mw and Tp values. Superimposed on each graph are prediction bands [13], which represent the level of condence on the model. For example, for a given Mw value, there is a 90% likelihood that the Tp value falls within the 90% prediction band that surrounds the regressed line. The large scatter observed in the data points in part can be attributed to the fact that the ground motions come from various events with various faulting mechanisms and geological conditions. The scatter translates into wide prediction bands, which indicate large uncertainties in predicting Tp from the regression equation. The magnitude of the earthquakes used in developing Equation (1) does not exceed 7.5 (Mw = 7:5 for the six simulated events, and Mw = 7:3 for the Landers earthquake). Somerville studied the magnitude 7.4 Izmit (Turkey, 1999) and magnitude 7.6 Chi-Chi (Taiwan, 1999) earthquakes [14]. He downplayed the prevalence of directivity e ects in the Chi-Chi earthquake and attributed the relatively small size of the forward directivity region to the shallow hypocenter, the small amount of up-dip slip, and the thrust-fault mechanism of this
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

702

B. ALAVI AND H. KRAWINKLER

Pulse Period - Magnitude Relationship


0.8 0.6 0.4 0.2 0 -0.2 6 6.2 6.4 6.6 6.8 7 7.2
log10Tp = - 1.76 + 0.31 Mw

Combined Ground Motions with Forward Directivity

log10Tp

data points regressed, mean 90% prediction band 80% prediction band 70% prediction band Taiwan, Tsaotun Turkey, Yarimca

7.4

7.6

7.8

Magnitude, Mw

Figure 15. Dependence of equivalent pulse period on magnitude.

earthquake [15]. Two of the pulse-type records used in his study are Yarimca (Turkey) and Tsaotun (Taiwan). The pulse period for the fault-normal component of these ground motions (from the location of a global peak in the velocity spectrum) is estimated to be 3:7 sec. and 4:2 sec:, respectively. Figure 15 compares these values with predictions from Equation (1), and indicates a good agreement. Adopting the functional form used in Equation (1), a similar study was conducted by Rodriguez-Marek [12], in which he investigated a larger set of recorded ground motions, spanning wider ranges of magnitude and distance. He classied the ground motions into two categories: (1) rock, including competent rock, weathered rock and shallow sti soil; and (2) soil, including deep sti clay and soft clay. This soil denition is broader than the one used in this study. He concluded that the existence of soil at the site elongates the pulse period, but the period elongation diminishes for large magnitudes. The conclusion from these studies is that there is a pattern of increase in pulse period with magnitude, but the scatter of the data is large, and regression results will depend on soil type, source mechanism, record selection, and on the procedure employed to dene the pulse period. What is needed most to improve predictions of Tp Mw relationships is the improvement of simulation techniques and the availability of more records. More research should take care of the former and time will take care of the latter. Somerville developed a model that expresses the PGV of the record as a function of moment magnitude and distance [2]. His proposed functional form is used here in a linear two-variable regression analysis that relates the equivalent pulse intensity parameter, ve , to Mw and R. A combination of ground motions, also used by Somerville, is used that includes the records listed in Table II and a subset of the simulated ground motions listed in Table III. This subset, which consists of simulations for Mw = 6:5 and 7.0 with strong forward directivity e ects (station f8), is most compatible with the recorded time histories [2]. Records with R values smaller than 3 km are not used in the derivation of the regression equation on account of the logarithmic form of this equation, which results in unreasonably large values for ve at small R values. The following equation results from the regression analysis: log10 ve = 2:03 + 0:65Mw 0:47 log10 R
Copyright ? 2004 John Wiley & Sons, Ltd.

(2)

Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

703

Effective Pulse Velocity


Data Points Surface Points

200 150 100 50 0 6 6.2 6.4 6.6 6.8 7 10 9 8 7 6 3 4 5

Veff (cm / sec)

R (km)

Mw

Figure 16. Dependence of equivalent pulse velocity on magnitude and distance.

Figure 16 illustrates this equation together with the data points to which the surface is tted. The solid circles identify the data points used in the regression analysis, and the empty circles represent points on the regression surface. The quality of the t appears to be relatively good. However, the data contain a relatively small number of recorded motions and a relatively large number of simulations, which is likely to bias the outcome of the regression analysis. The relatively steep slope of the surface at Mw = 7 translates into excessively large predictions for ve for events with Mw 7. PGV data from recent earthquakes (the 1999 Turkey and Taiwan earthquakes) provide a clear indication that PGV saturates at large magnitudes and that the predictions from Equation (2) are much too high for magnitudes greater than 7. When compared to Equation (2), the regression equation developed by Rodriguez-Marek [12] indicates a much weaker variation of PGV with magnitude, and for large magnitudes results in smaller pulse intensity. This is again an indication that Equation (2) cannot be considered a reliable measure for the pulse intensity of near-fault ground motions in very large magnitude earthquakes. In conclusion, it is claimed that the e ective velocity as dened in this paper (or in a simplied way, the PGV) together with the pulse period are good descriptors of the equivalent pulse. The functional forms of the regression equations relating Tp and ve to Mw and R are believed to be suitable, but further investigations and much more data are needed to develop more reliable regression coe cients and uncertainty measures. Nevertheless, Equations (1) and (2) are used in the following discussion to estimate the period and intensity of the equivalent pulse for specic Mw and R values. For the range of Mw and R values used in the discussion, the results are believed to be reasonable.

MAGNITUDE AND DISTANCE DEPENDENCE OF BASE SHEAR STRENGTH DEMANDS Given the moment magnitude, Mw , and the closest distance from the rupture plane, R, the equivalent pulse parameters Tp and ag; max (= ae = 4ve =Tp ) can be estimated from Equations
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

704

B. ALAVI AND H. KRAWINKLER

Figure 17. Preliminary results on magnitude and distance dependence of MDOF base shear strength demand spectra for constant ductilities, equivalent pulse P2: (a) Mw = 6, R = 10 km; (b) Mw = 6, R = 3 km; (c) Mw = 7, R = 10 km; and (d) Mw = 7, R = 3 km.

(1) and (2), and the pulse T=Tp curves presented in Figure 13 can be converted into T curves [ = Vy =W = (ag; max =g)]. The so obtained T curves represent base shear strength demand spectra for specied target ductilities, suitable for the types of frames represented by the generic frame used in this study. Examples of such strength demand spectra are presented in Figure 17 for various combinations of magnitude and distance, assuming that the ground motion is represented by pulse P2. Inherent is the assumption of the SRSS story shear strength distribution over the height of the structure. The graphs illustrate the magnitude and distance dependence of the base shear strength demands obtained from the equivalent pulse approach. To put these demands in perspective with present design practice, superimposed on each graph are two design strength curves related to the UBC 97 soil Type SD design spectra for Seismic Zone 4, one with and one without the code specied near-fault factor (denoted as UBC/4) [4]. The UBC near-fault factor is computed using Source Type A for Mw = 7 and Source Type C for Mw = 6. The design strength curves are obtained by scaling down the code base shear by a factor of 4, which is arrived at by choosing a response modication factor of 8 (close to that for special moment frames) and assuming a period independent overstrength factor of 2. These values are purely for comparison, as the real strength of a structure will vary on a case-by-case basis.
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

BEHAVIOR OF MOMENT-RESISTING FRAME STRUCTURES

705

A comparison of the near-fault strength demand spectra with the UBC= 4 curves reveals that for given magnitude and distance values, a structure designed according to present code provisions will experience quite di erent levels of inelasticity depending on the fundamental period. For instance, Figure 17(d) indicates that for Mw = 7, R = 3 km, and a ground motion of the type represented by pulse P2, structures designed according to the UBC 97 provisions (including the near-fault factor) may experience story ductility demands less than 4 when the period is longer than 3:4 sec:, but may experience ductility demands larger than 8 if the period is between 1.0 and 2:2 sec. Assuming that the = 4 spectrum for Mw = 7 and R = 3 km is representative of design conditions, it is observed that this spectrum has a much wider plateau (up to a period of about 2 sec.) than the UBC= 4 curve. This plateau is followed by a descending curve that falls below the UBC= 4 curve around 3:4 sec. This indicates that long-period and short-period structures are adequately protected by the UBC 97 provisions, but structures of intermediate period (from about 0.8 to 3:4 sec.) may experience ductility demands signicantly larger than 4.0. The introduction of the UBC near-fault factor is an improvement that gives recognition to the existence of the problem. However, the results of this study indicate that the code criteria do not provide a consistent level of protection against near-fault e ects, and in certain period ranges may provide inadequate protection. The problem cannot be solved by introducing additional factors to conventional design spectra, and a more rigorous approach appears to be necessary. The equivalent pulse concept can provide the basis for such an approach. CONCLUSIONS The study summarized in this paper addresses important response characteristics of SDOF and MDOF structures subjected to near-fault ground motions and basic pulses. Equivalent pulses, which are fully dened by a small number of parameters, are introduced to represent nearfault ground motions. Relationships are proposed for the magnitude and distance dependence of these parameters. A procedure for estimating the required design base shear strength is introduced based on the response characteristics of equivalent pulses. It is emphasized that the conclusions and results presented in this paper are applicable only within the context of the assumptions made. The conclusions of this study are summarized as follows:
Spectral values for individual near-fault ground motions can be several times the values For structures with T Tp , large elastic story shear forces are generated in the upper

given by the UBC 97 design spectrum over a wide range of periods.

stories of the structure.


These large elastic shear forces result in early yielding of upper stories when the structure

is relatively strong. When the strength is reduced, story ductility demands stabilize in the upper portion and the maximum demand migrates to the base, where it grows rather continuously with a further reduction in strength. For short-period structures (T 6Tp ) the maximum story ductility demands occur in the bottom portion regardless of strength. There are clear similarities between the response of frame structures to near-fault ground motions and the response to pulse-type excitations. Within the approximate period range of 0:375T=Tp 3:0, important response characteristics of near-fault ground motions can be represented by simple equivalent pulses.
Copyright ? 2004 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2004; 33:687706

706

B. ALAVI AND H. KRAWINKLER

Preliminary regression equations are developed that relate equivalent pulse parameters

to moment magnitude and closest distance to the fault rupture plane. Equations of this type, together with pulse MDOF strength demand spectra, can be used to estimate the required base shear strength for target story ductility ratios. For Mw = 7 and R = 3 km, a comparison between the base shear strength demand spectra obtained from the regression equations and UBC 97 design estimates indicates that code-compliant structures with an intermediate period may experience excessive ductility demands.
ACKNOWLEDGEMENTS

The research summarized in this paper was supported by a grant from the CUREE/Kajima Research Program, by the National Science Foundation through Grant CMS-9812478 of the USJapan Cooperative Research Program in Urban Hazard Mitigation, and by the California Department of Conservation as a SMIP 1997 Data Interpretation Project (Department of Conservation Contract No. 1097-601). This support is gratefully acknowledged. The constructive collaboration of Dr Paul Somerville on the ground motion aspects of this research is much appreciated.
REFERENCES 1. Somerville P, Smith N, Graves R, Abrahamson N. Modication of empirical strong ground motion attenuation relations to include the amplitude and duration e ects of rupture directivity. Seismological Research Letters 1997; 68:180 203. 2. Somerville P. Development of an improved ground motion representation for near fault ground motions. Proceedings of SMIP98 Seminar on Utilization of Strong-Motion Data, Oakland, CA, September 1998. 3. Structural Engineers Association of California. Recommended Lateral Force Requirements and Commentary. 7th Edition, SEAOC: Sacramento, California, 1999. 4. International Conference of Building O cials. Uniform Building Code. Whittier, California, 1997. 5. Somerville P, Smith N, Punyamurthula S, Sun J. Development of ground motion time histories for phase 2 of the FEMA/SAC steel project. Report No. SAC/BD-97-04, 1997. 6. Federal Emergency Management Agency. NEHRP Recommended Provisions for Seismic Regulations for New Buildings and Other Structures. 2000 Edition. FEMA 368. FEMA: Washington, DC, 2001. 7. Alavi B, Krawinkler H. E ects of near-fault ground motions on frame structures. John A. Blume Earthquake Engineering Center Report No. 138, Stanford University, February 2001. 8. Seneviratna G, Krawinkler H. Evaluation of inelastic MDOF e ects for seismic design. John A. Blume Earthquake Engineering Center Report No. 120, Stanford University, June 1997. 9. Chopra A. Dynamics of Structures: Theory and Applications to Earthquake Engineering. Prentice Hall: Upper Saddle River, New Jersey, 1995. 10. Stewart J, Chiou S, Bray J, Graves R, Somerville P, Abrahamson N. Ground motion evaluation procedures for performance-based design. PEER-2001/09, Pacic Earthquake Engineering Research Center, University of California, Berkeley, September 2001. 11. Sasani M, Bertero V. Importance of severe pulse-type ground motions in performance-based engineering: historical and critical review. 12th World Conference on Earthquake Engineering, New Zealand Society for Earthquake Engineering, Upper Hutt, New Zealand, 2000. 12. Rodriguez-Marek A. Near-fault site response. Ph.D. dissertation, Department of Civil and Environmental Engineering, University of California, Berkeley, Fall 2000. 13. Ramsey F, Schafer D. The Statistical Sleuth: a Course in Methods of Data Analysis. Duxbury Press: Belmont, California, 1996. 14. Somerville P. Characterization of near-fault ground motions. Proceedings of the USJapan Workshop on E ects of Near-Field Earthquake Shaking, San Francisco, CA, March 2000. 15. Somerville P. Magnitude scaling of near fault ground motions. Proceedings of the International Workshop on Annual Commemoration of Chi-Chi Earthquake, National Center for Research on Earthquake Engineering, Taipei, Taiwan, September 2000.

Copyright ? 2004 John Wiley & Sons, Ltd.

Earthquake Engng Struct. Dyn. 2004; 33:687706

Vous aimerez peut-être aussi