152
INTRODUCTION TO EARTHQUAKE ENGINEERING LECTURE NOTES Halûk Sucuoğlu Sinan Akkar Ankara 2011

Ce 490 Lecture Notes-combined

Embed Size (px)

Citation preview

Page 1: Ce 490 Lecture Notes-combined

INTRODUCTION TO EARTHQUAKE ENGINEERING

LECTURE NOTES

Halûk Sucuoğlu Sinan Akkar

Ankara 2011

Page 2: Ce 490 Lecture Notes-combined

2

Introduction to Earthquake Engineering

Contents

Chapter 1. Nature of Earthquakes 1.1 LIVING EARTH: CONTINENTAL DRIFT

1.1.1 Tectonic Plates 1.1.2 Plate Boundaries and Faults

1.2 FORMATION OF EARTHQUAKES 1.2.1 Fault Rupture 1.2.2 Types of Fault Motions 1.2.3 Elastic Rebound Theory

1.3 MAGNITUDE OF AN EARTHQUAKE 1.4 INTENSITY OF AN EARTHQUAKE

1.4.1 Observational Intensity 1.4.2 Instrumental Intensity: Strong Ground Motions

1.5 EFFECTS OF EARTHQUAKES ON STRUCTURES 1.4.1 Primary Effects: Ground Vibration 1.4.2 Secondary Effects: Fault rupture and Geotechnical Deformations

Chapter 2. Response of Simple Structures to Earthquake Ground Motions 2.1 INTRODUCTION 2.2 EQUATION OF MOTION: DIRECT EQUILIBRIUM 2.3 EQUATION OF MOTION FOR BASE EXCITATION 2.4 SOLUTION OF THE SDOF EQUATION OF MOTION

2.4.1 Free Vibration Response 2.4.2 Forced Vibration Response: Harmonic Excitation 2.4.3 Forced Vibration Response: Earthquake Excitation

2.5 EARTHQUAKE RESPONSE SPECTRA 2.5.1 Pseudo Velocity and Pseudo Acceleration Response Spectrum 2.5.2 Practical Implementation of Earthquake Response Spectra 2.5.3 Normalization of Earthquake Response Spectra

2.6 NONLINEAR SDOF SYSTEMS 2.6.1 Equation of Motion of a Nonlinear SDOF System 2.6.2 Nonlinear Force-Deformation Relations 2.6.3 Ductility and Strength Spectra for Nonlinear SDOF Systems

Page 3: Ce 490 Lecture Notes-combined

3

2.6.4 Ductility Reduction Factor (Rμ ) 2.6.5 Equal Displacement Rule

FURTHER READING Chapter 3. Response of Building Structures to Earthquake Ground Motions 3.1 INTRODUCTION 3.2 EQUATIONS OF MOTION 3.3 UNDAMPED FREE VIBRATION: EIGENVALUE ANALYSIS

3.3.1 Vibration Modes and Frequencies 3.3.2 Normalization of Modal Vectors 3.3.3 Orthogonality of Vibration Modes 3.3.4 Modal Expansion of Displacements

3.4 FORCED VIBRATION UNDER EARTHQUAKE EXCITATION: MODAL SUPERPOSITION

3.4.1 Modal Superposition Analysis Procedure 3.4.2 Response Spectrum Analysis 3.4.3 Equivalent Static (Effective) Modal Forces

FURTHER READING Chapter 4. Seismic Hazard Analysis

Chapter 5. Seismic Analysis Procedures in Earthquake Codes 5.1 INTRODUCTION 5.2 DESIGN GROUND MOTIONS: LINEAR ELASTIC RESPONSE SPECTRUM

5.2.1 Linear Elastic Acceleration Spectrum in the Turkish Code 5.2.2 Structure Importance Factor (I)

5.3 REDUCTION OF ELASTIC FORCES: DESIGN SPECTRUM 5.4 ANALYSIS PROCEDURES

5.4.1 Reduction of Degrees of Freedom: Static Condensation 5.4.2 Mode Superposition Procedure 5.4.3 Equivalent Static Lateral Load Procedure 5.4.4 Accidental Torsion

REFERENCES Chapter 6. Earthquake Resistant Design of RC Structures

Page 4: Ce 490 Lecture Notes-combined

4

2.

Response of Simple Structures to Earthquake Ground Motions

2.1 INTRODUCTION The main purpose of response analysis in earthquake engineering is the estimation of

earthquake induced forces and deformations in structures under the action of earthquake

ground motions. To this end, our approach will start with developing response analysis

procedures for the simplest dynamic system, called the single degree of freedom system.

These procedures are then extended to more complicated systems in the following chapters.

We will introduce the definition of a single degree of freedom system in this chapter

first, and then derive the equations of motion governing its free vibration response (simple

harmonic motion) and forced vibration response. Earthquake ground excitations lead to

special forms of forced vibration response.

Single Degree of Freedom (SDOF) System:

The deformed shape of the system at any instant can be represented in terms of a single

dynamic coordinate u(t), called the single degree of freedom.

A car with mass m connected to a fixed end with a spring with stiffness k, which is

free to move on rollers only in the lateral direction is a typical SDOF system (Fig. 2.1.a). An

inverted pendulum type of structure where the concentrated mass m is connected to the fixed

base with a massless cantilever column is also an ideal SDOF system (Fig. 2.1.b). In this case

the spring stiffness is identical to the lateral stiffness of the column, i.e. k=3EI/L3. The

motion of the mass and the elastic force which develops in the spring at any time t can be

represented by the dynamic displacement u(t), which is the single degree of freedom in both

systems. On the other hand, a pendulum where the mass m connected with a chord of length l

that swings about the fixed end of the chord in the gravity field g is another example of a

SDOF system (Fig. 2.1.c). In this case the degree of freedom is the angle of rotation θ.

Page 5: Ce 490 Lecture Notes-combined

5

(a) (b) (c)

Figure 2.1. Ideal SDOF systems: (a) a car on rollers, (b) an inverted pendulum structure, (c) a pendulum swinging in the gravity field.

More complicated dynamic systems can also be idealized as SDOF systems. Let’s

consider a cantilever column and a simple multistory frame in Fig. 2.2 where the lateral

deformation shapes exhibit variation along the height during motion. Both systems can be

defined as SDOF systems if the lateral dynamic deformation distribution u(x,t) along the

height x can be expressed as , φ where φ (x) is the normalized deformation

profile, i.e. φ 1. Lateral displacement at the top is the single degree of freedom.

Note that φ (x) is assumed to be constant and not changing with time. This is not

exactly correct, but practically acceptable. This assumption is valid for simple structural

systems. For example, φ (x)=(x/L)2 is an acceptable deformation shape for both SDOF

systems in Fig. 2.2 since it satisfies the boundary conditions u(0)=0 and u’(0)=0 at x=0.

(a) (b)

Figure 2.2. Idealized SDOF systems: (a) cantilever column, (b) multistory frame.

Page 6: Ce 490 Lecture Notes-combined

6

2.2 EQUATION OF MOTION: DIRECT EQUILIBRIUM

Consider two ideal SDOF systems in Fig. 2.3 with a mass, spring and damper. Damper

is the only difference between Figs. 2.2 and 2.3, which represents internal friction in the

actual mechanical system that is idealized as a SDOF system. Internal friction develops in

deforming mechanical systems such as shown in Fig. 2.2 due to rubbing of the molecules with

respect to each other during dynamic deformations. Internal friction leads to energy loss in a

vibrating system.

Figure 2.3. Ideal SDOF systems with mass m, stiffness k and damping c.

When the mass moves by a positive displacement u(t) with a positive velocity (t)

under an external force F(t), the spring develops a resisting force which is equal to k·u(t), and

the damper develops a resisting force which equals c· (t), both in the opposite directions. Free

body and kinetic diagrams of the masses in both SDOF systems are shown in Fig. 2.4.

Figure 2.4. Free body diagrams of the masses when they displace by t and moving with a velocity

of t and an acceleration of t at time t

Applying Newton’s second law of motion ΣF ma for dynamic equilibrium of the

mass in the lateral direction leads to

(2.1) or

(2.2)

This is a 2’nd order linear ordinary differential equation (ODE) with constant coefficients m, c

and k.

Page 7: Ce 490 Lecture Notes-combined

7

2.3 EQUATION OF MOTION FOR BASE EXCITATION

The base of the inverted pendulum moves with the ground during an earthquake

ground shaking with a ground displacement of ug (t) as shown in Fig. 2.5.a. There is no direct

external force F t acting on the mass when ground moves, but inertial force develops on the

mass according to Newton’s 2’nd law (F=ma) where a is the total acceleration of the mass

ütotal. It is the sum of ground acceleration and the acceleration of the mass relative to the

ground.

ütotal = üg + ü (2.3)

Free body diagram of the mass is shown in Fig. 2.5.b. Then, according to Newton’s 2’nd law,

ΣF mütotal yields

(2.4)

or ü ü 0 (2.5)

Transforming into the standard form of Eq.(2.2) leads to,

ü ü (2.6) where ü is considered as an effective force.

(a) (b)

Figure 2.5. (a) An SDOF system under base excitation, (b) free body diagram of the mass when it

displaces by t , while moving with a velocity of t and an acceleration of t at time t .

Page 8: Ce 490 Lecture Notes-combined

8

2.4 SOLUTION OF THE SDOF EQUATION OF MOTION The solution of a 2’nd order ODE can be obtained in two parts:

u(t) = uh(t) + up(t) (2.7)

where uh is the homogeneous solution and up is the particular solution. In vibration problem,

uh represents the free vibration response (F = 0) and up represents the forced vibration

response (F ≠ 0).

2.4.1 Free Vibration Response The motion is imparted by the initial conditions at t=0: u(0) = u0 (initial displacement) and

(0) = ν0 (initial velocity). The equation of free vibration is given by

ü 0 (2.8)

Dividing all terms by the mass m gives

ü 0 (2.9) Let 2ξ and . This is a simple replacement of the two coefficients and

in terms of two new coefficients ξ and , which have distinct physical meanings. The

dimensionless parameter ξ is the critical damping ratio, and ωn is the natural frequency

(rad/s). Vibration occurs only if ξ < 1. Then Eq. (2.9) can be written as,

ü ξ ü 0 (2.10)

Undamped Free Vibration (ξ = 0)

When damping is zero, Eq. (2.10) reduces to

ü 0 (2.11)

Eq. (2.11) represents simple harmonic motion. Only a harmonic function satisfies Eq. (2.11)

with a harmonic frequency of . Its most general form is a combination of sin and cos

functions with arbitrary amplitudes.

Page 9: Ce 490 Lecture Notes-combined

9

(2.12)

A and B are determined by introducing the initial conditions u(0) = u0 and (0) = ν0, leading

to

cos ν sin (2.13)

Eq. (2.13) is shown graphically in Fig. 2.6.

Figure 2.6. Undamped free vibrations of a SDOF system

Damped Free Vibration (0<ξ<1)

The presence of damping in free vibration imposes a decaying envelope on the

undamped free vibration cycles in Fig. 2.6. Decay is exponential, and decay rate depends on

, as given in Eq. (2.14).

ξ cos ν ξ sin (2.14)

The amplitude of harmonic vibration reduces exponentially at each cycle, and approaches

zero asymptotically as shown in Fig. 2.7.

Figure 2.7. Free vibrations of a damped (under-damped) SDOF system

Page 10: Ce 490 Lecture Notes-combined

10

Note that the term in brackets in Eq. (2.14) is similar to Eq. (2.13) where is

replaced by , which is the “damped” natural frequency given by Eq. (2.15).

1 ξ (2.15)

ξ ≤ 0.20 for structural systems in general, hence ωd ∼ ωn. Typical viscous damping ratios that

can be assigned to basic structural systems are given in Table 2.1.

Table 2.1. Typical damping ratios for basic structural systems

Structural type Damping ratio

Steel 0.02-0.03

Reinforced concrete 0.05-0.08

Masonry 0.10-0.20

Example 2.1. Consider the pendulum in Fig. (a) with mass m connected to a chord of length L, oscillating in the gravity field.

a) Determine its equation of motion. b) Solve the equation of motion for small oscillations θ when the motion starts with an

initial displacement θ0. a) At any θ(t), free body diagram of the mass is shown in Fig.(b), where T is the tension in the chord. Equation of motion in the t (tangential) direction can be written as

ΣF mat From Fig. (b),

‐mg sin θ mat mL Rearranging,

mL mg sin θ 0 (Solution a) (1)

(a) (b)

Page 11: Ce 490 Lecture Notes-combined

11

b) Eq. (1) is a 2nd order nonlinear ODE. Nonlinearity is due to the sin θ term. For small oscillations, sin θ ≈ θ. Hence, the equation of motion becomes linear. 0, or 0. With similitude to Eq. (2.11),

2 .

The solution from Eq. (2.12) is,

sin cos (2) Substituting θ (0) = θ0 and 0 0 into Eq. (2), we obtain cos√ (Solution b) Example 2.2. Determine the natural frequency of vibration for the system shown in Fig. (a) where the bar AB is rigid and it has no mass. The system in Fig. (a) is a SDOF system where the vertical displacement of end B can be employed as the DOF. The displacement variation of the SDOF system is always linear from A to B with a fixed shape as shown in Fig. (b). (a) (b) Since all forces are not directly acting on the mass, a direct formulation of the equation of motion is not possible. The conservation of energy principle provides a simpler approach.

(1)

where T is the kinetic energy and U is the potential energy at any time t, given by

(2)

Here, we should consider from Fig. (b) that the velocity of the mass in terms of the DOF u is

. Hence, . Substituting into Eq. (1), and taking time derivative of both sides,

2 2 0 or

Page 12: Ce 490 Lecture Notes-combined

12

4 0 Accordingly,

2 (Solution)

Example 2.3. A single story, single bay portal frame is given below in Fig. (a). a) Determine the equation of free vibration and the natural period of free vibration, b) Determine the equation of motion under base excitation üg(t).

a) The portal frame is a SDOF system with the fixed deflection shape shown in Fig. (b). The lateral displacement u of the mass m at the roof is the degree of freedom. Free body diagram of the roof mass is given in Fig. (c). (a) (b) (c) Applying Newton’s second law of motion ΣF ma for dynamic equilibrium of the mass in the lateral direction leads to

(1) Rearranging,

150

is the effective stiffness of the portal frame, and m is the mass. Accordingly,

2 √ (Solution a)

b) Eq. (1) can be written for base excitation as,

15

or, (Solution b)

Page 13: Ce 490 Lecture Notes-combined

13

2.4.2 Forced Vibration Response: Harmonic Excitation Harmonic excitation can either be applied as an external harmonic force, or an

effective harmonic force due to a harmonic base excitation ü sin . Equation of

motion under harmonic excitation can then be written as

ü ü sin (2.16)

The homogeneous solution is identical to the damped free vibration response in Eq.

(2.14), where vibration occurs at the free vibration frequency .

ξ (2.17)

A1 and A2 are the arbitrary amplitudes that have to be determined from the initial conditions

at t=0. However the initial conditions are imposed on the general (total) solution, not on the

homogeneous solution alone.

The particular solution is assumed to be composed of sin and cos functions where

vibration occurs at the forced vibration frequency .

(2.18)

The arbitrary harmonic amplitudes G1 and G2 are determined by using the method of

undetermined coefficients for up.

ξ

; ξξ

; (2.19)

General Solution The general solution is the combination of homogeneous and particular solutions from

Eqs. (2.17) and (2.18), respectively. Substituting G1 and G2 from Eq. (2.19) into Eq. (2.18),

simplifying and collecting into u(t)=uh(t)+up(t), we obtain

ξ cos ξ

ξ (2.20)

A1 and A2 are determined from the initial conditions as indicated above.

Page 14: Ce 490 Lecture Notes-combined

14

In damped systems under harmonic excitation, uh is called the transient and up is called

the steady-state response since the transient part decays with time as shown in Fig. 2.7. If the

transient part is ignored, then the remaining component up can also be expressed as

(2.21)

where

ξ

, ξ (2.22)

Here, ρ is the amplitude, and θ is the phase lag (delay) between up and . It can be

shown by expanding that Eq. (2.21) with Eq. (2.22) is identical to the second

(steady-state) term in Eq. (2.20). The variation of ρ with the frequency ratio β and the

damping ratio ξ is plotted in Fig. 2.8, which is called the frequency response function. It can

be observed that the response displacement amplitude amplifies as β approaches to unity

whereas increase in damping ratio reduces the level of amplification.

Figure 2.8. Frequency response function for damped systems under harmonic excitation

Resonance

When β =1, i.e. the forcing frequency is equal to the natural frequency of vibration

in Eq. (2.19), Eq. (2.18) reduces to

/

ξ ξ 1 cos (2.23)

Eq. (2.23) is plotted in Fig. 2.9.a. The amplitude of displacement cycles increase at every

cycle and asymptotically approach /ξ

.

Page 15: Ce 490 Lecture Notes-combined

15

Meanwhile, if ξ→ 0, L’Hospital rule gives

/ cos (2.24)

The second term in the parenthesis indicates a linear increase of displacement amplitude with

time, without any bound. Eq. (2.24) is plotted in Fig. 2.9.b.

Eqs. (2.23) and (2.24) define a vibration phenomenon called the resonance. In

mechanical systems, resonance causes very high displacements which usually lead to

collapse.

(a) (b)

Figure 2.9. Resonance in (a) damped, (b) undamped SDOF systems under harmonic excitation

2.4.3 Forced Vibration Response: Earthquake Excitation A SDOF system under earthquake ground acceleration is shown in Fig. 2.10. The

excitation function F t or ‐müg t can rarely be expressed by an analytical function in the

case of earthquake ground excitation. Ground acceleration üg t is given numerically.

Closed-form analytical solution similar to Eq. (2.20) is not possible. Then the solution is

obtained by using numerical integration techniques. The most practical and also the most

popular method is the step-by-step direct integration of the equation of motion (Newmark,

1956).

Figure 2.10. An SDOF system under earthquake excitation

Page 16: Ce 490 Lecture Notes-combined

16

Numerical Evaluation of Dynamic Response Consider the equation of motion of a SDOF system at time t = ti and t = ti+1 ≡ ti + Δt

where Δt is small.

ü (2.25.a)

ü ü (2.25.b)

Subtracting (2.25.a) from (2.25.b) gives,

ü ü (2.26)

or

∆ü ∆ ∆ ∆ (2.27) where

∆ · · · (2.28)

Eq. (2.27) contains three unknowns ∆ , ∆ , ∆ü . Therefore it is indeterminate.

However, we may impose two kinematical relations between these three response parameters,

such as ü and . This can be achieved by assuming a variation of

acceleration ü(t) over Δt, then integrating twice to calculate (t) and u(t) within Δt. We

assume that ü , and at the beginning of the time step are known from the previous step.

Two common assumption can be made on the variation of ü(t) over Δt : constant

average acceleration and linear acceleration variation.

Constant Average Acceleration Consider the variation of acceleration ü(t) over a time step Δt shown in Fig. 2.11.

This actual variation can be estimated by an approximate, constant average acceleration

variation given in Eq. (2.29).

Fig. 2.11. Actual and estimated acceleration variations over a time step Δt.

Page 17: Ce 490 Lecture Notes-combined

17

ü ü ü ; 0 ∆ (2.29)

It should be noted here that the actual variation of acceleration on the left hand side of Eq.

(2.29) is not known, and hence üi+1 on the right hand side is also an unknown. This is merely a

transfer of unknown from a function to a discrete value by the assumption of constant average

acceleration variation.

We can integrate constant acceleration variation given in Eq. (2.29) twice, in order to

obtain the variations of velocity and displacement over the time step Δt, respectively. This

process is schematized in Fig. 2.12.

Fig. 2.12. Integration of constant average acceleration variation over the time step Δt

The first integration is from acceleration to velocity, i.e. the integration of d =üdτ.

ü (2.30)

Page 18: Ce 490 Lecture Notes-combined

18

Substituting ü from Eq. (2.29) into Eq. (2.30) and integrating, we obtain

ü ü (2.31)

Eq. (2.31) can also be written for τ=ti+1 at the end of the time step which gives

ü ü (2.32) Then, substituting from Eq. (2.31) into and integrating over Δt,

∆ (2.33) we obtain

∆ ∆ ü ü (2.34) The terms ui 1, and üi 1 at ti+1 in Eqs. (2.32) and (2.34) are the unknowns.

Let ü ü ü ü 2ü ∆ü 2ü . When this identity is substituted

into Eqs. (2.32) and (2.34) and rearranged, two new equations are obtained:

∆ ∆ ∆ü 2ü (2.35)

∆ ∆ ∆ ∆ü 2ü (2.36)

∆ , ∆ and ∆ü are the new three unknowns in Eqs. (2.35) and (2.36). Combining these

equations with Eq. (2.27) forms a system of three coupled linear equations with three

unknowns, and can be solved through elimination.

Lets rearrange (2.35) and (2.36) to express Δüi and ∆ in terms of Δui. From Eq. (2.36),

∆ü ∆

∆ ∆

2ü (2.37) Substituting ∆ü above into Eq. (2.35),

∆ ∆

∆ 2 (2.38) Finally, substituting Δüi and ∆ from Eqs. (2.37) and (2.38) into Eq. (2.27) and rearranging,

we obtain

∆ ∆ (2.39) where

(2.40) is the instantaneous dynamic stiffness, and

Page 19: Ce 490 Lecture Notes-combined

19

∆ ∆ ∆

2 2 ü (2.41) is the effective dynamic incremental force. Note that in Eq. (2.40), i.e. dynamic

stiffness does not change at each time step i.

The recursive solution starts at i = 0 with 0 0 and 0 0 as the initial

conditions. This procedure is unconditionally stable: Errors do not grow up with the recursion

steps. However ∆ 10 is required for accuracy.

The step-by-step direct integration procedure described above is formulated as an

algorithm below, which can be easily coded with conventional software (FORTRAN,

MathLab, Excel, etc.).

Integration Algorithm

1. DEFINE m, c, k, 0 0, 0 0, Fi F ti and Δt 2. ü 3. CALCULATE k* from Eq. (2.40) 4. 1 5. CALCULATE ∆ from Eq. (2.41)

6. ∆ ∆

7. CALCULATE ∆ and ∆ü from Eqs. (2.37) and (2.38) 8. · · ∆ · for · , , ü 9. GO TO 4

2.5 EARTHQUAKE RESPONSE SPECTRA

Consider various SDOF systems with different T, but the same ξ, subjected to a

ground excitation as shown in Fig. 2.13. Note that T1 < T2 < T3 <··· in Fig. 2.13.

2

2

Fig. 2.13. Different SDOF systems under earthquake ground excitation

t

gu.. Düzce NS

17/08/1999

T1 ξ T2 ξ T3 ξ T4 ξ

u1(t)

u2(t)

u3(t)

u4(t)

Page 20: Ce 490 Lecture Notes-combined

20

We can calculate the displacement response of each SDOF system u(t) by direct

integration. Time variations u(t) and ü(t) of 5 percent damped SDOF systems with T1=0.5 s,

T2=1.0 s and T3=2.0 s under the NS component of 1999 Düzce ground motion are plotted in

Fig. 2.14.

Fig. 2.14. Time variations of displacement and acceleration responses of several SDOF systems under the NS component of 1999 Düzce ground motion

We can select the peak displacement response from each u(t) function, and define this

value as the spectral displacement Sd, where,

| | (2.42) Since each u(t) is a function of T and ξ, Sd also varies with T and ξ. Hence,

, (2.43) Similarly, spectral acceleration can be defined as the peak value of total acceleration

ü ü (2.44) where

, (2.45) Sa and Sd values are marked on Fig. 2.14 for the response of each SDOF system.

Page 21: Ce 490 Lecture Notes-combined

21

Accordingly, Sd and Sa in Eqs. (2.43) and (2.45) can be plotted as functions of T and ξ.

When this process is repeated for a set of damping ratios, a family of Sa and Sd curves are

obtained. The family of these curves is called the acceleration response spectra and

displacement response spectra of an earthquake ground motion, respectively. Acceleration

and displacement response spectra of the NS component of 1999 Düzce ground motion are

plotted in Fig. 2.15. The peak values indicated in Fig. 2.14 are also marked on Fig. 2.15.

The duration effect is almost lost in the spectral information since an earthquake

response spectrum only considers the time when peak response occurs. This is practical for

design, however a long duration ground motion may cause low cycle fatigue and consequent

degradation. We cannot obtain such detailed information from a response spectrum.

Fig. 2.15. Acceleration and displacement response spectra of the NS component of 1999 Düzce ground motion

It can be observed from Fig. 2.15 that when T=0, Sa =üg,max (PGA) and Sd = 0. On the

other hand, when T approaches infinity, Sa approaches zero and Sd approaches ug,max (PGD).

These limiting situations can be explained with the aid of Fig. 2.16.

T=0 is equivalent to ωn = ∞, i.e. the system is infinitely stiff. Then the spring does not

deform (u=0, hence Sd = 0) and the motion of the mass becomes identical to the motion of

ground. Accordingly, maximum acceleration of the mass becomes identical to the acceleration

of the ground which makes their maximum values equal.

T approaches infinity when ωn approaches zero. Hence the system becomes infinitely

flexible. An infinitely flexible system has no stiffness and it cannot transmit any internal

lateral force from the ground to the mass above. Ground moves while the mass stays

stationary during the earthquake. The total displacement of the mass is zero (utotal=u+ug=0).

Accordingly |u|max=|ug|max, or Sd =PGD. Similarly, the total acceleration of the mass is zero

(ütotal=ü+üg =0) which makes Sa =0 from Eq. (2.44).

Page 22: Ce 490 Lecture Notes-combined

22

Fig. 2.16. Response of infinitely stiff and infinitely flexible SDOF systems to ground excitation

2.5.1 Pseudo Velocity and Pseudo Acceleration Response Spectrum

Pseudo spectral velocity PSv and pseudo spectral acceleration PSa can be directly

obtained from Sd. Their definitions are given in Eqs. (2.46) and (2.48) below. PSv and PSa are

very close to Sv and Sa respectively, for ξ< 0.20.

Pseudo Velocity

, ξ . , ξ (2.46)

for ξ < 0.20. It is related to the maximum strain energy Es stored in the

SDOF system during the earthquake.

, (2.47) Pseudo Acceleration

, ξ , ξ (2.48) for ξ<0.20. It is related to the maximum base shear force at the support of

the SDOF system during the earthquake.

Lets consider an undamped SDOF system under a ground excitation üg. Its equation of

motion is,

ü ü 0 (2.49) Therefore,

ü ü | | (2.50)

Page 23: Ce 490 Lecture Notes-combined

23

Then, or (2.51)

for ξ=0. Comparison of Eqs. (2.48) and (2.51) indicates that PSa=Sa when ξ=0.

The shear (restoring) force and base shear force which develop in a SDOF system

during an earthquake ground excitation are shown in Fig. 2.17. Base shear force becomes

maximum when the relative displacement is maximum, i.e.,

, (2.52) Now, let’s consider a damped SDOF system. Eq. (2.52) is replaced by

, (2.53) Hence, the maximum base shear force in an undamped SDOF system can directly be obtained

from PSa through Newton’s second law.

Fig. 2.17. Internal shear force and base shear force developing in an SDOF system under earthquake ground excitation

2.5.2 Practical Implementation of Earthquake Response Spectra

If Sd, PSv and PSa are available for an earthquake ground excitation, then we can easily

obtain the maximum values of response displacement, strain energy, internal elastic force and

base shear force of an SDOF system from these spectral charts. The only data we need for a

SDOF system is its natural vibration period Tn and viscous damping ratio ξ.

The maximum base shear force in Eq. (2.53) can be formulated as follows:

, . (2.54)

Page 24: Ce 490 Lecture Notes-combined

24

where W is the weight in the gravity field. Then,

, (2.55)

The ratio of maximum base shear force to weight is called the base shear coefficient,

which is a practical yet very important parameter in earthquake engineering analysis and

design. It is denoted with C, and it can be obtained directly from PSa.

, (2.56)

Example 2.4. Consider the portal frame in Example 2.3. The properties assigned are, column size: 0.40x0.50 m, E=250,000 kN/m2, L=3m and m=25 tons. Determine the maximum displacement of the roof if the frame is subjected to the 1999 Düzce NS ground motion. Viscous damping ratio is 5 %. The natural vibration period was defined in Example 2.3. When the numerical values given above are substituted, Tn =1.3 seconds is calculated. Then the spectral acceleration can be determined from Fig. (2.15) at T=1.3 seconds as Sa = 4 m/s2.

The effective force acting at the roof mass is,

Feff mSa = 25 tons x 4 m/s2 = 100 kN

The effective stiffness expression was also derived in Example 2.3. When the numerical values given above are substituted, keff = 578.7 kN/m is calculated. Finally,

0.173 m (Solution)

2.5.3 Normalization of Earthquake Response Spectra

Spectral acceleration shapes for different earthquake ground motions exhibit

significant variations, as shown in Fig. 2.18.a for a suit of 10 ground motions. They are

usually normalized by selecting a fixed damping ratio first, usually ξ = 0.05, then removing

the effect of with peak ground acceleration üg,max (PGA) by dividing Sa(T) with PGA for all T.

The spectral acceleration shapes normalized with respect to PGA in Fig. 2.18.b. display less

variation compared to the non-normalized spectra in Fig. 2.18.a.

Page 25: Ce 490 Lecture Notes-combined

25

Fig. 2.18. (a) Acceleration response spectra of 10 ground motions, (b) acceleration response spectra of 10 ground motions normalized with respect to their peak ground accelerations.

It is possible to obtain statistical averages of Sa T / PGA over period (Fig. 2.18.b).

However, this is usually done first by grouping the ground motions with respect to the soil

conditions of the recording stations (sites), then obtaining the mean values of Sa T / PGA

values over T. This exercise was first carried out by Seed et al. (1976) which indicate the

effect of local soil conditions on the shape of mean acceleration spectra (Fig. 2.19). These

shapes form the basis of earthquake design spectra defined for local soil conditions, which

will be discussed in Chapter 4.

(b)

(a)

Page 26: Ce 490 Lecture Notes-combined

26

Fig. 2.19. (a) Mean normalized acceleration response spectra of ground motions classified with respect to local soil conditions (Seed, H. B., Ugas, C. and Lysmer, J. (1976) , “Site-dependent spectra for earthquake-

resistant design”, Bull.Seism. Soc. Am. 66, No.1, 221-243). 2.6 NONLINEAR SDOF SYSTEMS The lateral forces which act on linear elastic structural systems under severe

earthquake ground motions are usually very large. It can be observed from Fig. 2.18.a that the

effective lateral forces (m·PSa) are at the order of the weight (mg) in the period range of 0.4-

1.0 seconds. Fundamental vibration periods of most of the building structures fall into this

period range. Designing structures for such high levels of lateral forces is not economical and

feasible for a very seldom event such as a strong earthquake which may occur only with a

small probability during its service life. The preferred approach in seismic design is to provide

a lateral strength Fy that is less than the elastic strength demand Fe, however implement a

plastic deformation capacity to the system such that it can deform beyond the linear elastic

range during a strong ground motion.

When the yield displacement capacity of the lateral load resisting system is exceeded,

the slope of the restoring force-deformation curve, or stiffness softens. A typical force-

deformation path of a SDOF system subjected to a single cycle of large ground displacement

is shown in Fig. 2.20. This type of nonlinear behavior is called material nonlinearity in

mechanical systems because softening occurs due to the deterioration of material properties at

large displacements, similar to the stress-strain behavior of steel and concrete materials. Fs is

the restoring force (internal resistance), Fy is the yield force capacity and uy is the yielding

displacement in Fig. 2.20.

Page 27: Ce 490 Lecture Notes-combined

27

Fig. 2.20. Variation of internal force Fs with displacement u along a nonlinear force-deformation path. 2.6.1 Equation of Motion of a Nonlinear SDOF System Equation of motion of a nonlinear SDOF system is also mathematically nonlinear.

ü ü (2.57)

The restoring force term Fs creates the nonlinearity in the equation of motion since Fs (u) is a

nonlinear function. In a linear system, Fs is equal to k u which is a linear relationship between

Fs and u whereas Fs = Fs (u) in a nonlinear system which implies that the tangent stiffness k is

not constant as in a linear system, but changes with the displacement u. The variation of Fs

with u along a nonlinear force-deformation path is schematized in Fig. 2.20.

The equation of motion can be written in incremental form with the aid of Fig. 2.21.

∆ü ∆ ∆ ∆ ∆ü (2.58)

The incremental variation of restoring force Fs with u can be estimated with

∆ ∆ (2.59) where

∆ and ∆ (2.60) and ki is the tangent stiffness at ui.

Page 28: Ce 490 Lecture Notes-combined

28

Fig. 2.21. Incremental variation of Fs and u.

Step-by-step direct integration algorithm developed for linear elastic systems in 2.4.3

can be applied to nonlinear systems if tangent stiffness ki (ui) is known and hence can be

updated at each load step i. This requires a priori knowledge of the Fs (u) function, which is

called the hysteresis relationship.

2.6.2 Nonlinear Force-Deformation Relations Hysteresis relations are composed of a set of rules by which the variation of Fs is

defined in terms of the variation history of u during the previous loading steps. This is called a

hysteresis model. Two basic hysteresis models are employed in earthquake engineering:

Elasto-plastic and stiffness degrading models. The elasto-plastic model is usually employed

for representing the hysteretic flexural behavior of steel structures whereas the stiffness

degrading model represents the hysteretic flexural behavior of concrete structures

respectively, under loading reversals induced by an earthquake ground motion.

The set of rules which define elasto-plastic and stiffness degrading hysteretic behavior

with strain hardening are shown in Fig. 2.22. Fy and uy are the yield strength and yield

displacement respectively, k is the initial elastic stiffness and αk is the strain hardening

stiffness after yielding where α is usually less than 10 percent. When α=0, the system is

elastic-perfectly plastic.

Page 29: Ce 490 Lecture Notes-combined

29

Elasto-Plastic Model Stiffness Degrading Model (steel members) (RC members)

Fig. 2.22. Elasto-plastic and stiffness degrading hysteresis models Elasto-Plastic Model

There are two stiffness states in the elasto-plastic model (Fig. 2.22.a): k or αk. Initial

loading (0-1 or 0-1’) starts with the stiffness k, and when the internal resistance reaches the

plastic state, the system yields and deforms along the yield plateau with post-yield stiffness αk

(1-2). Unloading and reloading (2-3; 4-5; 6-7) take place along the elastic paths with the

initial stiffness k. When the direction of loading changes from loading to unloading or vice

versa along these paths, the stiffness does not change. On the other hand, when the internal

resistance reaches the plastic state along these paths at points (3, 5 or 7), plastic deformations

occur along the yield plateau with post-yield stiffness αk (3-4 or 5-6).

Stiffness Degrading Model

In the stiffness degrading model (Fig. 2.22.b), unloading and reloading stiffnesses are

different. Unloading from the yield plateau takes place with the initial elastic stiffness k (2-3

or 5-6). Reloading then follows with a degraded stiffness defined from the point of complete

unloading (3, 6 or 8) to the maximum deformation point in the same direction which occurred

during the previous cycles (points 4, 2 or 9). Unloading from a reloading branch before

reaching the yield plateau also takes place with the stiffness k (7-8).

Under an earthquake excitation, nonlinear systems can only develop a resistance

bounded by their lateral yield strength Fy, but they respond at larger displacements. Consider

Page 30: Ce 490 Lecture Notes-combined

30

three SDOF systems with the same initial stiffness k and period T, subjected to the same

earthquake ground motion üg (Fig. 2.23). Their properties, from the weakest to strongest are:

System 1: Elasto-plastic with yield strength Fy1 and yield displacement uy1

System 2: Elasto-plastic with yield strength Fy2 and yield displacement uy2 (Fy2 > Fy1)

System 3: Linear elastic, i.e. Fy = ∞

Fig. 2.23. Three SDOF systems with the same initial stiffness k, but different yield strengths Fyi, subjected to a strong ground excitation cycle.

We can expect that the weakest system (System 1) deforms most, to an absolute

maximum displacement of umax1, while system 2 deforms to a lesser maximum displacement

of umax2. Meanwhile the linear elastic system deforms to a maximum displacement of ue under

the same earthquake ground motion. These maximum absolute displacements umax1, umax2 and

ue are called the displacement demands of the earthquake from systems 1, 2 and 3,

respectively. We can also define these demands in terms of a dimensionless deformation ratio

μ, called the ductility ratio. , (2.61)

Therefore, an earthquake ground motion demands more ductility from systems with less

strength.

⇒ (2.62) Fe is the elastic force demand above, and μe = 1 theoretically for linear elastic systems.

The force terms on the left in Eq. (2.61) are the strengths (capacities) whereas the

terms on the right are the ductility ratios (demands).

Page 31: Ce 490 Lecture Notes-combined

31

2.6.3 Ductility and Strength Spectra for Nonlinear SDOF Systems

We can solve the nonlinear equation of motion

ü ü (2.63) for different elasto-plastic systems with the same Fy and ξ, but different ki (or Ti ) under the

same earthquake ground motion üg. Accordingly, we can obtain the maximum displacement

umax,i corresponding to each system with ki , or Ti. This is schematized in Fig. 2.24 under a

strong ground excitation cycle.

Fig. 2.24. Three SDOF systems with the same strength Fy, but different initial stiffnesses ki, subjected to a strong ground excitation cycle.

Then,

versus Ti can be plotted as a spectrum, called the ductility spectrum.

If this process is repeated for different Fy values, we obtain ductility spectra for constant

strength values Fy. The Fy values are usually expressed as a ratio of mg in these charts.

Ductility spectra obtained for the Düzce 1999 ground motion is shown in Fig. 2.25.

Fig. 2.25. Ductility spectra of the 1999 Düzce NS ground motion for different constant strength ratios.

Page 32: Ce 490 Lecture Notes-combined

32

Next, we can convert the μ - T (ductility) spectra to a Fy – T (strength) spectra by

graphical interpolation. If we assume a constant ductility value on Fig. 2.25, it intersects each

Fy curve at a different T value. Hence, a set of Fy-T values are obtained for a constant ductility

ratio of μ. We can plot the set of Fy-T values for constant ductility as the strength spectrum.

When this process is repeated for different constant ductility values, a family of constant

ductility curves is obtained which is called the strength spectra. The strength spectra obtained

for the Düzce 1999 ground motion is shown in Fig. 2.26. This graphics is also called the

inelastic acceleration spectra (Sai – T) for constant ductility μ.

Fig. 2.26. Strength spectra or inelastic acceleration spectra of the 1999 Düzce NS ground motion for

different constant ductility ratios.

The correspondence between the ductility spectra in Fig. 2.25 and the strength spectra

in Fig. 2.26 can be explained with a simple example. Point A in Fig. 2.25 is on the Fy =0.40

mg curve at T= 0.21 s with μ=2. Similarly, point B is on the Fy =0.20 mg curve at T= 0.71 s

with μ=4. These two points are also marked on Fig. 2.26, at the same period values on the

corresponding constant ductility curves for μ=2 and 4. Their Fy values are exactly the same.

If we know the period T and strength Fy of our SDOF system, then we can directly

calculate the ductility demand of the earthquake from the inelastic acceleration spectra. On

the other hand, if we have a given (estimated) ductility capacity μ for our system, then we

can determine the required (minimum) strength for not exceeding this ductility capacity under

the considered ground motion. This is very suitable for the force-based seismic design.

Page 33: Ce 490 Lecture Notes-combined

33

2.6.4 Ductility Reduction Factor (Rμ )

Rμ is defined as the ratio of elastic force demand Fe to the yield capacity Fy of the

nonlinear SDOF system with the same initial stiffness k, under the same earthquake ground

excitation üg (Eq. 2.64), which is depicted in Fig. 2.27.

(2.64)

Fig. 2.27. Force-displacement responses of a linear elastic and a nonlinear SDOF system under an earthquake ground excitation üg.

There is a corresponding ductility demand from the nonlinear system with

the initial stiffness k=Fy/uy and the initial period . If this process is repeated for

several nonlinear systems with different Fy and T values as in Fig. 2.23, a set of

curves can be obtained, which can be plotted as a spectra. The spectra obtained

for the Düzce 1999 and El Centro 1940 ground motions are shown in Fig. 2.28.

Usually it is observed that oscillates around when where Tc is called

the corner period of the ground motion. Then when . This assumption is

valid for the mean spectrum of many ground motions, although it is a crude

approximation for a single ground motion.

Page 34: Ce 490 Lecture Notes-combined

34

Fig. 2.28. spectra of the 1999 Düzce NS and 1940 El Centro NS ground motion components for different constant ductility ratios.

The mean spectrum can be idealized in a simple form shown in Fig. 2.29.

1 1 Corner period

Fig. 2.29. Idealized form of the spectrum.

By using an exact spectrum, or the idealized one, we can obtain the

inelastic acceleration spectrum Sai and inelastic displacement spectrum Sdi from the

corresponding linear elastic acceleration and displacement spectrum, Sae and Sde ,

respectively. If we exploit the identities

// and /

we obtain

and . (2.65. a, b)

0

1

2

3

4

5

6

7

8

9

0 0.5 1 1.5 2 2.5 3 3.5

T (sec)

R

μ = 1

μ = 2

μ = 4

μ = 8

Düzce NS 17/08/1999; ξ = 5%

0

2

4

6

8

10

12

14

0 0.5 1 1.5 2 2.5 3 3.5

T (sec)

R

μ = 1μ = 2

μ = 4

μ = 8

El Centro NS 19/05/1940; ξ = 5%

Page 35: Ce 490 Lecture Notes-combined

35

Eq. (2.65.a) can be employed for obtaining the inelastic acceleration response spectra

Sai directly from the linear elastic acceleration spectrum Sae by selecting an ratio. This is

very practical for seismic design since factors for different types of structural systems are

defined in seismic design codes. Similarly, inelastic displacement response spectra Sdi can be

obtained from the linear elastic displacement response spectrum Sde by using Eq. (2.65.b).

Inelastic acceleration (yield acceleration) spectra and inelastic displacement spectra calculated

for the Düzce 1999 ground motion by employing Eqs. (2.65) are shown in Figs. 2.30.a and

2.30.b respectively for several factors.

It is noteworthy to compare the strength spectra in Fig. 2.26 with the yield acceleration

spectra in Fig. 2.30.a. In an elasto-plastic system, strength and yield acceleration are related

through Fy=may. However the inelastic spectral curves are slightly different at short periods. If

was not constant for each curve in Fig. 2.30.a but it was a function of period as in Fig.

2.28.a, then the two spectra would be the same.

Fig. 2.30.b implies that the response displacements of linear elastic and inelastic

SDOF systems are very close. This property is discussed further in the following section.

Fig. 2.30. Inelastic acceleration (yield acceleration) and displacement response spectra of the 1999 Düzce NS ground motion component for different Rμ factors.

(a)

(b)

Page 36: Ce 490 Lecture Notes-combined

36

2.6.5 Equal Displacement Rule

For medium and long period SDOF systems (T > 0.5 second), Rμ  =  μ implies the

“equal displacement rule”, which is derived in Eq. (2.66) below with the aid of Fig. 2.31, Eqs.

(2.61) and (2.64).

⇒ ⇒ (2.66)

Fig. 2.31. Force-displacement relationships of linear elastic and nonlinear SDOF systems under a ground excitation. ue ≈ umax implies the equal displacement rule.

Equal displacement rule can be simply tested by plotting the variation of inelastic

displacement ratio umax / ue with T which is shown in Fig. 2.32 for the ground motions

employed in Fig. 2.18. The rule is verified for T > 0.1 second since the mean umax / ue ratio

approaches unity after this period.

Fig. 2.32. Variation of inelastic-to-elastic maximum displacement ratio with period.

Page 37: Ce 490 Lecture Notes-combined

37

Equal displacement rule is employed as a practical tool in displacement-based

assessment and design. When it can be assumed that the equal displacement rule holds for

ground motions in general, the inelastic displacements of a yielding system can directly be

obtained from the linear elastic displacement response spectrum of its simpler counterpart.

FURTHER READING Hibbeler, R. C., Engineering Mechanics- Dynamics, Prentice Hall, Englewood Cliffs, N. J., 2001, Chapter 22. Beer, F. P. and Johnston, E. R., Vector Mechanics for Engineers- Dynamics, McGraw-Hill, 1982, Chapter 19. Chopra, A. K., Dynamics of Structures, Prentice Hall, Englewood Cliffs, N. J., 2002, Chapters 1-3, 5-7. Humar, J. L., Dynamics of Structures, Prentice Hall, Englewood Cliffs, N. J., 1990, Chapter 8.

Page 38: Ce 490 Lecture Notes-combined

38

3. Response of Building Structures to

Earthquake Ground Motions

3.1 INTRODUCTION

Building structures are multi degree of freedom (MDOF) systems where more than

one displacement coordinate is necessary for defining the position of the system during

motion.

The minimum number of displacement coordinates required to define the deflected

shape of the system properly at any time “t ” during motion is the number of “degrees of

freedom”, DOF. They are the independent coordinates (displacement u t , rotation θ t , etc.)

that change with time.

3.2 EQUATIONS OF MOTION We will develop the equation of motion of a MDOF system by employing a shear

frame for brevity. A shear frame is a single-bay, N-story frame consisting of flexible columns

and fully rigid girders where the story masses are assigned to the girders. An N-story shear

frame under base excitation üg (t) is shown in Fig. 3.1. Since the girders are rigid, there are no

joint rotations at the joints and transverse displacements at both ends of a girder are identical.

Accordingly only one degree of freedom ui is sufficient for each story i, which is along the

girder where the mass mi is assigned. Each story i has a total shear stiffness ki that is

composed of the column shear stiffnesses 12EI/ h3 in that story.

The equation of motion under base excitation has the same form with Eq. (2.5), where

the scalar displacement variables , and for a SDOF system are replaced with the

vectorial displacement variables , and for a MDOF system. Similarly, the scalar mass,

stiffness and damping property terms are replaced with the associated matrix quantities:

Page 39: Ce 490 Lecture Notes-combined

39

ü ü ü 0 (3.1) where

ü ü . and

11

1

(3.2)

The vector l is transmitting the ground displacement ug to the story DOF’s above as rigid

body displacements. It is called the influence vector. l = 1 for shear frames since a unit

displacement at the ground is transmitted equally to all DOF’s defined at the stories above.

Fig. 3.1. An N-story shear frame subjected to the ground excitation üg (t)

When ütotal and l are substituted from Eq. (3.2) into Eq. (3.1), we obtain

ü ü (3.3) where

00 0

0

0 0

(3.4)

is the mass matrix, and

(3.5)

is the stiffness matrix. Each stiffness coefficient ki in Eq. (3.5) represents the total lateral

stiffness of the i’th story that is composed of the column shear stiffnesses as indicated above.

Page 40: Ce 490 Lecture Notes-combined

40

∑ 12 (3.6)

The displacement vector is composed of the N lateral story displacements (degrees of

freedom).

(3.7)

It should be noted that the mass matrix in Eq. (3.4) is a lumped matrix which is

indicating no coupling between the story masses. Moreover, the stiffness matrix in Eq. (3.5) is

tri-diagonal, hence the lateral stiffness of a story is coupled with the lateral stiffnesses of the

story below and above only (close-coupling). These are inherent properties of the shear frame

in Fig. 3.1.

There is no analytical method for obtaining the coefficients of the damping matrix c in

Eq. (3.3) from the damping properties of structural members. There is a practical approach for

obtaining the damping matrix of a MDOF system, called Rayleigh damping (Chopra 2001,

Chapter 11). The construction of damping matrix is not required however in the following

approach.

3.3 UNDAMPED FREE VIBRATION: EIGENVALUE ANALYSIS When the force term on the right-hand-side of Eq. (3.3) is zero and the damping is

ignored, we obtain the undamped free vibration equation:

ü 0 (3.8)

Free vibration can be induced by the initial conditions at t=0.

0 , 0 (3.9)

If we can impose a “special” initial shape u0, then we observe harmonic free vibration

(simple harmonic motion) with a fixed displacement profile along the height. A fixed profile

indicates fixed proportionality of the story displacements with respect to each other. Vibration

with a fixed displacement profile is identical to a single degree of freedom response which

was previously discussed in Section 2.1 and shown in Fig. 3.2.b and c. These special

displacement profiles are the “natural mode shapes”, and their corresponding harmonic

Page 41: Ce 490 Lecture Notes-combined

41

vibration frequencies are the “natural frequencies of vibration”. There are N such mode

shapes for an N-DOF system. Typical displacement profiles representing such mode shapes

are illustrated in Figs. 3.2.b and 3.2.c for a three story shear frame. When free vibration is

induced with a non-special, or general initial displacement shape, then the profile of this

initial shape cannot be retained during free vibrations, and a modality does not develop as

shown in Fig. 3.2.d.

We have to carry out eigenvalue analysis for determining the natural mode shapes and

natural vibration frequencies.

(a) (b) (c) (d)

Fig. 3.2. A 3-story shear frame in free vibration. (a) Shear frame properties, (b) and (c) harmonic free vibrations with special initial shapes, (d) non-harmonic free vibration where the initial shape

degenerates. Numbers on top indicate time sequence of deflections.

3.3.1 Vibration Modes and Frequencies

At a given mode, the displacement vector varies harmonically with time whereas its

shape profile remains “fixed”. Then, we can express the modal displacement vector as the

product of a harmonic function of time and a shape function.

· (3.10)

Page 42: Ce 490 Lecture Notes-combined

42

Here, describes the displacement profile along the height, or the mode shape whereas

is the time dependent amplitude of this profile. Their product in Eq. (3.10) gives the

modal displacement shape for a mode n at any time t during free vibration. This assumption is

analogous to the method of “separation of variables” in solving partial differential equations.

Since free vibration motion with a mode shape is harmonic, we can assume a

harmonic function for .

cos (3.11)

When is substituted from Eq. (3.11) into Eq. (3.10) and the displacement vector in Eq.

(3.10) is differentiated twice with respect to time, an expression for acceleration vector is

obtained.

· · (3.12)

In Eq. (3.12), and are the modal vibration frequency (eigenvalue) and mode shape

(eigenvector) of the n’th mode, which have to be determined through an inverse solution

approach.

Substituting u and ü from Eqs. (3.10) and (3.12) respectively into the equation of free

vibration motion (3.8), we obtain

0 (3.13)

Here, = 0 is not an acceptable solution for Eq. (3.13), because it implies no vibration

(trivial solution). Therefore,

0 (3.14.a) or

( · 0 (3.14.b)

This is a set of N-homogeneous algebraic equations. = 0 is also a trivial solution

(no deformation) for Eq. (3.14). A non-trivial solution is possible only if the determinant of

is zero (Cramer’s Rule):

0 (3.15)

Page 43: Ce 490 Lecture Notes-combined

43

Eq. (3.15) is equivalent to an Nth order algebraic equation with N roots. The values are the

roots, or the eigenvalues (n=1, 2, …. N). If is known for a mode n, then we can determine

the corresponding shape vector from Eq. (3.14).

Summary

For an N-DOF structural system, there are N pairs of eigenvalues and eigenvectors , ,

n=1, 2, ….. N. Their values are related to the mass and stiffness properties of the system. The

system can vibrate in a simple harmonic motion independently at each mode, with the profile

at the associated angular frequency .

Example 3.1. A 2DOF system is given in Figures (a) and (b), which are dynamically identical. Determine its eigenvalues and eigenvectors.

(a) (b) The equation of motion for free vibration, from Eq. (3.8), can be written as

00

üü 0

0 (1)

Then Eq. (3.15) is applied to the given problem.

det

0 (2)

A closed form solution cannot be determined from Eq. (2). This is possible however if we make a simplification in the parameters. Let , and . Then Eq. (3) can be obtained from Eq. (2), which is called the characteristic equation.

3 0 (3) Eq. (3) is a quadratic algebraic equation in where n=1, 2. The roots of the quadratic equation are and , which are the eigenvalues (note that the roots are not . Solution of Eq. (3) yields the following roots:

Page 44: Ce 490 Lecture Notes-combined

44

√ 0.618 (4)

√ 1.618 (5)

The eigenvectors will be determined from Eq. (3.14). For the given problem,

2 0

0 ; 1, 2 (6)

From row 1; 2 0 (7)

From row 2 : 0 (8)

We cannot find a unique solution for and from Eqs. (7) and (8) because they are a set of homogeneous equations. We can rather express in terms of for both n=1 and n=2.

(9) Let = 1 in Eq. (9) for n=1 and n=2. Then we first substitute into Eq. (9) and determine 1.618. Next, we substitute into Eq. (9) and determine

0.618. Accordingly, the modal vectors, or the eigenvectors for the two modes are determined.

1.01.618 ; 1.0

0.618 (10) The mode shapes for the 2DOF system in Fig. (b) are plotted below.

Example 3.2. Identify the degrees of freedom of the system given in Figure (a). Then determine its eigenvalues and eigenvectors.

Page 45: Ce 490 Lecture Notes-combined

45

Solution The frame has 2 DOF’s, which are defined at the top end A of the cantilever column and shown in Fig. (b). Although this is correct and consistent for static analysis, we have to transfer these DOF’s to point B for dynamic analysis since the point mass is assigned to the B end of the rigid girder. The original (uA , θA) and the transferred (new) degrees of freedom (uB, vB) are shown in Fig. (b). Note that these two sets of DOF’s are dependent since uB=uA (rigid body translation of AB) and vB =l  θA (rigid body rotation of AB). The kinematic relation between (uB , vB) and (uA , θA) is sketched in Fig. (c). The stiffness equations for the first and second set of DOF’s can be written as (determine as an exercise),

4 and

(1.a, b)

Then the mass and stiffness matrices for the second set of DOF’s are,

00

(2.a, b)

Let’s assume l=h for simplicity. Then 0 gives,

(3) In Eq. (3), . The two roots of the quadratic Eq. (3) can be determined as

0.789 15.211 Substituting and into Eq. (3.14) and solving the homogeneous set of linear equations, we obtain the two eigenvectors.

n=1: 1; 1.869

n=2: 1; 0.535 The mode shapes are sketched below.

Page 46: Ce 490 Lecture Notes-combined

46

3.3.2 Normalization of Modal Vectors

Let where is called the nth modal mass. If we divide

by ; then

1 (3.16)

The modal vector is now normalized with respect to the modal mass . This is

practical in numerical applications because it reduces the amount of arithmetical

computations. Most of the computational software directly calculates mass normalized modal

vectors in earthquake engineering practice.

Example 3.3. Consider Example 3.1. Calculate the mass normalized modal vectors.

1.0 1.618 00 1.0

1.618 3.618

1.0 0.618 00

1.00.618 1.382

The mass parameter m can be neglected in both terms since m is arbitrary. Then,

√ . 1.0

1.618 0.5260.851 and

√ . 1.0

0.618 0.8510.576

are the normalized modal vectors. It can be verified that the normalized modal vectors satisfy

1 and 1

Hence, 1 for mass-normalized modal vectors.

3.3.3 Orthogonality of Vibration Modes

Let’s consider two modes n and m, with , and , . From Eq. (3.14.a),

and (3.17. a, b) Pre-multiplying Eq. (3.15.a) with

and Eq. (3.17.b) with respectively,

and (3.18. a, b)

Page 47: Ce 490 Lecture Notes-combined

47

Now, transposing both sides of Eq. (3.18.b) and considering that and

due to the symmetry of both matrices,

(3.19) Finally, subtracting Eq. (3.19) from Eq. (3.18.a),

0 (3.20)

Since in general,

0 (3.21)

This is the condition of orthogonality of modal vectors with respect to the mass matrix. A

similar orthogonality condition with respect to the stiffness matrix follows from Eq. (3.19).

0 (3.22)

Therefore modal vectors are orthogonal with respect to m and k.

Example 3.4. Consider Example 3.1. Verify orthogonality of modal vectors with respect to the mass matrix.

1.0 1.618 00 1.0

0.618 0 !

A similar orthogonality condition can also be verified for Example 3.2. 3.3.4 Modal Expansion of Displacements

Any displacement vector can be expressed as a linear combination of the

orthogonal modal vectors :

… … . . N (3.23)

in Eq. (3.23) are the modal amplitudes, or modal coordinates.

For a set of , we can determine by employing the orthogonality property of

modal vectors. Let’s pre-multiply all terms in Eq. (3.22) by .

N N (3.24)

Page 48: Ce 490 Lecture Notes-combined

48

All parentheses terms are zero due to modal orthogonality with respect to mass, except the

term. Then,

(3.25)

The denominator term is equal to 1 if the modes are mass normalized.

Example 3.5. Determine the modal expansion of 11 in terms of the modal vectors

determined in Example 3.4.

0.526 0.851 1 00 1 1

1 1.377

0.325

Substituting into Eq. (3.23),

1.377 0.526

0.851 0325 0.8510.526 1.0

1.0

3.4 FORCED VIBRATION UNDER EARTHQUAKE EXCITATION:

MODAL SUPERPOSITION

We will reconsider the equation of motion of a MDOF system that was given by Eq.

(3.3).

ü ü (3.3)

u can be expanded in terms of modal vectors by using Eq. (3.23).

(3.26)

Substituting u(t) from Eq. (3.26) into Eq. (3.3) and calculating the appropriate time

derivatives in Eq. (3.3), we obtain

ü (3.27)

Page 49: Ce 490 Lecture Notes-combined

49

Pre-multiply each term in Eq. (3.27) by ,

ü (3.28)

Only those terms with r=n are non-zero due to the orthogonality of modes. Although this is

theoretically valid for and , we can also assume the orthogonality of modal vectors with

respect to .

ü (3.29)

The terms in the first, second and third parentheses on the left hand side are the modal

mass , modal damping and modal stiffness , respectively. The term on the

right hand side is called the modal excitation factor, .

(3.30)

(3.31)

(3.32)

(3.33)

When the parentheses terms in Eq. (3.29) are replaced with the definitions in Eqs.

(3.30) to (3.33), a compact form is obtained.

(3.34)

Dividing all terms by and introducing the modal damping ratio and modal vibration

frequency from Section 2.4.1 leads to a final normalized form.

2ξ ü (3.35)

Eq. (3.35) is valid for all modes, n = 1, 2… N. This is equivalent to a SDOF system in the

modal coordinate .

Page 50: Ce 490 Lecture Notes-combined

50

Equations (3.29-3.35) describe the modal superposition procedure where the system

of N-coupled equations of motion of the MDOF system in Eq. (3.3) is replaced with N-

uncoupled equations of motion of equivalent SDOF systems in Eq. (3.35). This procedure

provides significant advantages because working with coupled stiffness and mass matrices is

much more difficult in applying numerical integration methods compared to integrating the

uncoupled equations of motion separately.

Now, let’s recall the equation of motion of a SDOF system under base excitation

ü from Eq. (2.6). When Eq. (2.6) is normalized similar to the normalization of Eq. (3.34) into

Eq. (3.35), we obtain

ü 2ξ ü (3.36)

The only difference between Eq. (3.35) and Eq. (3.36) is the term applied to the ground

excitation ü in the modal equation of motion. Therefore the solution procedures developed

for SDOF systems under earthquake excitation in Chapter 2 are also valid for solving Eq.

(3.35).

3.4.1 Modal Superposition Analysis Procedure Modal superposition analysis procedure for solving Eq. (3.3) is summarized below in a

stepwise form.

1. Carry out eigenvalue analysis of Eq. (3.8) and determine the modal properties

, ) for n=1, 2……N.

2. Construct Eq. (3.34) or (3.35) for each mode n.

3. Solve Eq. (3.34) by using the methods developed for SDOF systems in Chapter 2 (ü

is scaled by ), and determine for n=1, 2……N.

4. Transform from modal to physical coordinates by using Eq. (3.26).

Page 51: Ce 490 Lecture Notes-combined

51

3.4.2 Response Spectrum Analysis

The third step in the mode superposition analysis procedure above can also be

performed by response spectrum analysis in a very simple manner. Let Sd T, ξ be the

displacement spectrum for üg(t). Then,

, ξ

(3.37)

and

, . (3.38)

Also

,

(3.39)

where (3.40)

Accordingly,

, , . (3.41)

In Eq. (3.41) above, , is the maximum value of the n’th mode displacement

vector . However since the time dependence is lost in Eq. (3.44), we cannot apply Eq.

(3.26) directly and combine the maximum modal displacement for obtaining the maximum

displacement distribution . Since all , and accordingly , do not occur

simultaneously,

, , … … … N,

or

: : . . . . :

Modal responses are independent from each other, and the maximum values of a

response parameter (displacement, rotation, internal force, moment, etc) occur at different

times at each mode, without any synchronization. Hence a statistical combination is necessary

Page 52: Ce 490 Lecture Notes-combined

52

for obtaining the maximum combined response. The SRSS (Square Root of the Sum of

Squares) rule provides good approximation for combining the modal maxima of displacement

components.

, , , … . . N ,

/ (3.42) or, in general

, , , … . . N ,/ (3.43)

SRSS is equally applicable for estimating the maximum value of any force parameter

(moment, shear, stress, etc.) or displacement parameter (curvature, rotation, displacement,

strain, etc.) from the superposition of the associated maximum modal values.

Example 3.6. Determine the maximum displacement distribution of the 3-story shear frame in Fig. (a) under the acceleration spectrum given in Fig. (b). The results of eigenvalue analysis are also given below. (a) (b)

140 000 / 175 000 k : total lateral stiffness for both columns

0.3140.6861.00

0.500.50

1.00

1.000.686

0.313

15.84 / 34.64 / 50.50 /

0.40 0.18 0.125

When we enter the response spectrum with the modal period values, we determine the modal spectral acceleration values.

1.0

Then the modal masses and modal excitation factors are determined.

Page 53: Ce 490 Lecture Notes-combined

53

φ 374 700 350 000 531 000

1 525 350 175 000 165 550

1.40 0.50 0.31

Let , (drop the max index). are obtained from Eq. (3.38) and maximum modal displacements are determined from Eq. (3.41).

5.47 0.41 0.12

1.723.765.47

0.200.20

0.41

0.120.10

0.04

Finally, the modal spectral displacements are combined with the SRSS rule for obtaining the maximum story displacement distribution.

1.72 0.20 0.12 1.743.76 0.20 0.10 3.77

5.47 0.41 0.04 5.50

It may be noted that , i.e. the first mode displacements dominate the total displacement distribution. In particular,

~ √5.47 0.41 0.04 = 5.50 cm

3.4.3 Equivalent Static (Effective) Modal Forces

We can define an equivalent static force vector for each mode n, which produces the

modal spectral displacements when they are applied to the MDOF system.

Page 54: Ce 490 Lecture Notes-combined

54

At any time t during dynamic response, dynamic equilibrium requires

(3.45)

Substituting from Eq. (3.10) into Eq. (3.11),

(3.46) We can express the modal forces in a simpler form. Eq (3.14) for free vibration can be written

as

(3.47) Multiplying each term by gives

(3.48)

Substituting the right hand side of Eq. (3.48) for the middle term in Eq. (3.46), we obtain

(3.49) This is a more practical expression since the diagonal matrix m is easier to work with

compared to the banded matrix k having off-diagonal terms.

If we employ response spectrum analysis, then , in Eq. (3.49). Then,

substituting , from Eq. (3.38) into Eq. (3.49),

(3.50)

Finally, we obtain a simplified expression for the modal spectral force vector after rearranging

Eq. (3.50).

(3.51) The total force at the base (modal base shear force) is equal to 1 where 1 is the unit

vector.

1 1 1 (3.52)

Then,

Page 55: Ce 490 Lecture Notes-combined

55

(3.53)

where is the effective modal mass. With this definition, the spectral response at

each mode under an earthquake base excitation that is expressed by its acceleration response

spectrum can be represented on a simple sketch of an equivalent SDOF system as shown in

Fig. (3.3).

Fig. 3.3. SDOF representation of spectral modal response at mode n.

Effective modal mass has an important practical aspect such that the sum of effective

modal masses for all modes is equal to the total mass of the building system.

(3.54)

This is exact for a shear frame, and quite accurate for an actual building structure. Effective

modal mass can be directly calculated from the mass matrix and the n’th mode vector.

1 (3.55)

Page 56: Ce 490 Lecture Notes-combined

56

Example 3.7. Calculate the modal force vectors for the frame in Example 3.6. Also calculate the effective modal masses, modal base shear forces, and modal moments at the top end of the first story columns. Combine these forces and moments by SRSS for calculating the total base shear force and first story column top moment.

Modal Forces

1.402 0 00 2 00 0

0.3140.6871.00

.

1 5093 3022 408

858858858

1 064728167

Effective modal masses

From Eq. (3.55),

736,550 kg 84.2% 87,500 kg 10% 51,275 kg 5.8%

~ 875 000 100%

M is the total mass where M=5m=5x175,000 kg=875,000 kg. Therefore the sum of effective modal masses is equal to the total mass (inaccuracy is due to the decimal truncation in modal vectors). Modal base shear forces

736,550 9.81 7226 ; 858 ; 503

Note that ∑ for all n =1-3.

~ 7294

Modal column moments

where V is the shear force in the column and h is the story height. at the first story columns. Then the modal column top moments at the first story are;

12

12 · 7226 1806 ; 214.5 ; 126

~ 1823 ·

Page 57: Ce 490 Lecture Notes-combined

57

Example 3.8. Consider the frame given in Example 3.2. If the frame is subjected to the ground excitation defined by the acceleration spectrum given in Example 3.6, determine the displacement of the mass at end B, the base shear force and base moment at the support, and the top moment of the column. Let EI=4000 kN-m2, m=4 tons and h=4 m. Solution Tn = 2π/ωn, which gives T1 = 1.786 s and T2 = 0.407 s from the results of Example 3.2. The corresponding spectral accelerations can be determined from the acceleration response spectrum of Example 3.6, as Sa1= 0.34 g and Sa2=1.0 g.

; 17.99 tons 5.145 tons 1 ; 3.48 tons 6.14 tons

(1)

0.193 1.193

Modal amplitudes

; 0.0522 m 0.0492 m (2) Modal displacement vectors

; 0.05220.0976 0.0492

0.0263 (3)

where ; 1, 2

Displacements at the B end (SRSS combination)

0.0522 0.0492 0.0717 m (Both modes contribute)

0.0976 0.0263 0.1011 m (1’st mode dominant)

Internal forces in column OA

Let’s denote the bottom end (fixed end) of the column by O. The end forces (lateral force and bending moment) of column OA can be determined both by the stiffness analysis of the column by using the modal end displacements, or by applying the equivalent static modal forces and calculating the associated modal internal forces. We will do both. a) Stiffness analysis: Let’s write the stiffness equation for column OA at the n’th mode.

4 2

2 4

(4)

Page 58: Ce 490 Lecture Notes-combined

58

Note that 0 (fixed end), and . Inserting the associated modal displacements from Eq. (3), together with EI and h values into Eq. (4),

2.55 kN 29.5 kN. m2.55 kN

19.3 kN. m

and 46.76 kN

86.95 kN. m 46.76 kN

100.1 kN. m

Base shear force and base moment at O, and moment at A (SRSS combination)

2.55 46.76 46.83 kN (2’nd mode dominant)

29.5 86.95 91.82 kN. m (2’nd mode dominant)

19.3 100.1 101.94 kN. m (2’nd mode dominant)

b) Equivalent static modal forces: From Eq. (3.51),

. ; 2.57 4.81 N

and 46.81 25.04 N

The modal forces and the associated modal moment diagrams of the column OA are shown on the frame below.

Mode 1 Mode 2

Modal base shear forces, base moments and moments at A can be calculated from statics. Then their SRSS combinations give the final values.

2.57, 46.81 , 2.57 46.81 46.88 kN

29.5, 87.08 , 29.5 87.08 91.94 kN. m

19.3, 100.1 , 19.3 100.1 102.04 kN. m

These values are very close to the values calculated from stiffness analysis. The differences are due to truncation errors.

Page 59: Ce 490 Lecture Notes-combined

59

FURTHER READING Chopra, A. K., Dynamics of Structures, Prentice Hall, 2001, Chapters 10-13. Clough, R. W. and Penzien, J., Dynamics of Structures, McGraw-Hill, Inc., 1993, Chapters 9-12.

Page 60: Ce 490 Lecture Notes-combined

60

4.

Seismic Hazard Analysis 4.1 INTRODUCTION Seismic hazard analysis (SHA) defines the level of ground-motion intensity (e.g.,

PGA, PGV, spectral ordinates Sa at different vibration periods, etc.) for design ground

motions. Seismic hazard is one of the fundamental components of seismic risk in earthquake

prone regions. When seismic hazard (ground motion intensity) is combined with seismic

vulnerability (weakness or proneness of the constructed environment to earthquake effects),

seismic risk is obtained.

The methodology implemented in SHA primarily considers the fundamental features

of the seismic sources that would trigger the hazard in the vicinity of construction area. The

information gathered from the seismic characteristics of the sources is further elaborated to

estimate the ground-motion intensity which is specific to the purposes of the engineering

project.

This chapter describes the essentials of fundamental techniques implemented in SHA

and summarizes their key components. Seismic hazard analysis is divided into two broad

branches where the subject is treated either in a deterministic, or a probabilistic manner.

The first two sections describe the common components in deterministic and

probabilistic SHA (abbreviated as DSHA and PSHA, respectively). The latter two sections

present DSHA and PSHA separately, placing particular emphasis on their conceptual

differences. In the last section, the basic code approach for the definition of design spectrum

is introduced.

4.2 SEISMICITY AND EARTHQUAKE RECURRENCE MODELS The first step in SHA is to study the seismic sources and past seismic events in the

area of construction. Seismic sources (fault lines and area sources) are mainly described in

geological and seismotectonic studies. Seismic activity (i.e., annual occurrence rate of

Page 61: Ce 490 Lecture Notes-combined

61

earthquakes) is investigated from the historical and contemporary seismic event catalogs.

Historical event catalogs are specific studies and are available as published reports, books or

scientific articles. The contemporary seismic event catalogs can be compiled from several

seismological agencies (e.g., International Seismological Center, http://www.isc.ac.uk/). The

survey on the seismic catalogs results in the location and size of earthquakes of engineering

interest (i.e., events larger than a certain magnitude level) in the study area. Fig. 4.1 presents

sample maps that show the location of active faults and spatial distribution of seismic events

in the area of interest for a specific seismic hazard study.

Fig. 4.1. A sample illustration that shows the fault lines or sources (black solid lines in the top panel) and spatial distribution of seismic events in the vicinity of construction site (green flag

in both panels).

The information extracted from above studies serves for identifying the maximum

event and the computation of closest source-to-site distance that are used in DSHA (Section

4.4). The temporal distribution of earthquakes in the area of interest defines the earthquake

Page 62: Ce 490 Lecture Notes-combined

62

recurrence model that is used in PSHA (Section 4.5). Given a threshold magnitude level (i.e.,

Mthreshold) the earthquake recurrence model gives the frequency of events for M ≥ Mthreshold.

This logarithmic relationship expressed in Eq.(1) is proposed by Gutenberg and Richter in

1954 by studying the earthquakes in Southern California.

(1)

Equation (1) describes the frequency of earthquakes as a power of magnitude (power-

law). Therefore, exponential probability distribution would describe the probability of

occurrence rate of earthquakes. The parameter a in Eq. (1) represents the number of events

per unit time above M = 0, which describes the seismic activity rate in the area of interest.

(Large a implies higher seismic activity). The slope term b defines the ratio of small and large

magnitude events. A steep slope means the dominancy of smaller magnitude events with

respect to larger magnitudes. In contrast, smaller b values advocate a higher contribution of

large magnitude events to the overall seismic activity in the region.

Calculating the magnitude-recurrence model as suggested by Richter is a

straightforward procedure after compiling the seismic event catalogs for the area under

consideration. After filtering the aftershocks (to warrant independency between seismic

events) the compiled data are sorted in ascending magnitude order. The cumulative number of

events (N) of magnitude M or greater is found and it is normalized by the time span covered

by the catalog to compute the annual rate of earthquakes exceeding a threshold magnitude

level (n). A straight line fit to the plot of log10(n) vs. M yields Gutenberg-Richter magnitude

recurrence model. This procedure is illustrated in Fig. 4.2 by using a sample case. The table

on the left lists the compiled catalog information with number of events equal or above certain

magnitude values (N). The magnitudes in the example are moment magnitudes. When

compiling the earthquake catalogs, the magnitude scale of the events should be the same in

order to describe the size of each earthquake consistently. The total catalog duration is 102

years in this example and the annual frequencies are determined by normalizing N with the

observation duration. The semi-log plot in the right panel shows the actual variation of annual

exceedance rates (black circles) as a function of M. This panel also shows the least squares

straight line fit to the empirical data. The straight line fit shown in Fig. 4.2 suggests that the

annual frequency of earthquakes is ~0.025 for an earthquake of magnitude 6.

Page 63: Ce 490 Lecture Notes-combined

63

Fig. 4.2. Application of Gutenberg-Richter magnitude recurrence model on a real data set.

The logarithmic relationship (exponential distribution) given in Eq. (1) can

equivalently be represented as in Eq. (2) where ν0 = 10a and β = b ln(10). Note that now ν0

and β represent the activity and slope terms.

(2)

Each seismic source can generate a certain range of earthquakes in size. This is clearly shown

in the physical problem discussed in Fig. 4.2. Therefore, often a lower magnitude bound (i.e.,

Mmin) other than the theoretically introduced 0 is implemented, and a modified form of Eq. (2)

is used in the hazard computations. The modified form of Eq. (2) is presented in Eq. (3).

(3)

The activity term, νM,min, in the new formulation refers to the annual frequency (or annual

exceedance rate) of earthquakes above Mmin. The minimum magnitude bound is generally

defined from the catalog information (for example, it is approximately M = 4 in the catalog

given in Fig. 4.2) or from the engineering objectives of site-specific hazard study. Many

M# of events of magnitude M or greater (N)

N normalized by catalogue duration (n)

4.05 95 0.934.15 87 0.854.25 77 0.754.35 74 0.734.45 63 0.624.55 57 0.564.65 50 0.494.75 44 0.434.85 33 0.324.95 22 0.225.05 20 0.205.15 17 0.175.25 14 0.145.35 10 0.105.45 8 0.085.65 5 0.055.95 2 0.026.15 1 0.01

log10(n) = 3.91-0.93M

0.001

0.010

0.100

1.000

10.000

4 5 6

annu

al e

xcee

danc

era

te, n

Magnitude

Page 64: Ce 490 Lecture Notes-combined

64

engineers consider the magnitude 5 event as the smallest damaging earthquake and use this

magnitude to cap the minimum magnitude level.

A careful consideration of Fig. 4.2 would suggest that the straight line fit deviates

from the actual variation of data (black circles) at the large magnitude end. The data tend to

be truncated towards larger magnitudes. This observation can be explained again by the

ability of seismic sources generating a certain magnitude range of earthquakes. As shown in

Fig. 4.2, the maximum event produced by the considered source has a magnitude of ~ 6.2. On

the other hand the straight line model, due to its inherent nature, cannot cap the observed

maximum magnitude (i.e., Mmax) in the source. In other words, the straight line model will

keep calculating smaller and smaller annual frequencies for magnitudes larger than the

observed maximum magnitude in the source. Such kind of deficiencies can be prevailed by

using more rigorous magnitude-annual frequency relationships to fine tune the magnitude

recurrence models. The observed saturation towards larger magnitudes can be addressed

better by using the truncated exponential distribution. This model caps the annual frequency

of earthquakes for a given Mmax. Moreover, maximum likelihood (ML) regression technique

would better describe the overall variation of the data because least-squares regression

procedure gives higher weight to the outliers that result in discrepancies between the actual

data and fitted model.

Figure 4.3 shows the magnitude recurrence model for the same data in Fig. 4.2 by

using ML regression technique and truncated exponential distribution. As explained in the

previous paragraph truncated exponential distribution requires the limiting maximum

magnitude. Mmax is determined either from the compiled catalog information or using the

empirical equations that consider the fault dimensions for the estimation of Mmax. As a general

approach the analyst adds 0.5 magnitude units to the observed Mmax in the compiled catalog to

consider the uncertainty in the magnitude of future events that can occur in the area of

interest. In passing, for the example given in Figure 4.3, none of the above approaches are

followed and Mmax is taken as 6.2, which is determined directly from the compiled catalog

(see the table in Fig. 4.2).

Page 65: Ce 490 Lecture Notes-combined

65

Fig. 4.3. Same data presented in Fig. 4.2 but the fitted distribution is truncated-exponential

with the ML regression technique. (The dashed arrows are discussed in the latter parts of this section).

Exponential distribution discussed in the previous paragraphs approximate the

earthquake recurrence in area sources. Due to their finite geometry, the individual faults tend

to produce earthquakes within a specific magnitude range. These are called as “characteristic”

magnitudes. Therefore the magnitude recurrence of individual faults is generally described by

the “characteristic model.” This model assumes an exponential relationship for the annual

frequency of small magnitude events. Larger magnitude earthquakes (designated as

characteristic events) are supposed to occur more frequently in the characteristic model and

they are assumed to vary uniformly along a certain large magnitude range. This way the

model acknowledges the higher occurrence probability of characteristic events. The width of

the characteristic part of the model can be 0.5 magnitude units. The application of

characteristic model requires a proper estimation of slip rates on individual faults as well as

Mmin and Mmax. Figure 4.4 presents a typical example from this model.

0.001

0.010

0.100

1.000

10.000

4 5 6

annu

al e

xcee

danc

era

te, n

Magnitude

Page 66: Ce 490 Lecture Notes-combined

66

Fig. 4.4. Characteristic magnitude-recurrence model (modified from McGuire, 2004).

The reciprocal of annual exceedance rates of magnitude recurrence models describe

the recurrence interval of events for a specified magnitude. For example, the recurrence

interval of M 6.1 event is 100 years (= 1/0.01) in the example given in Fig. 4.3 (red dashed

arrows). In other words, the average time between the occurrences of an event with M = 6.1 is

100 years for the given case.

The overall discussions up to this point address that the magnitude recurrence models

describe the occurrence of earthquakes above a certain magnitude level in time. The potential

occurrence or recurrence of an earthquake above a certain magnitude level in a sequence of

events can be seen as a Bernoulli sequence problem. In the Bernoulli sequence, the sequence

of events is called as trials and occurrence of earthquake above a given magnitude is the

successful event. In earthquake related problems, the earthquakes can occur at any instant and

it may occur more than once at a given time. In such a case, the probability of occurrence of

the event is appropriately modeled with a Poisson process. The Poisson process is stationary

in time and has no memory on the occurrence of earthquakes. This implies equal chances of

occurrence between the consecutive events and their independency from each other. These

features do not represent the actual mechanism behind the occurrence of earthquakes because

stresses relieved in one segment of the source due to the occurrence of an earthquake results

in the accumulation of stresses in another segment that increases the rupture potential in that

segment. Nonetheless the Poisson process is frequently used in converting the occurrence

frequencies of earthquakes to a probability for certain time interval t. This time interval is

called as the exposure time and it is generally considered as the nominal life time of major

structures.

Ann

ual e

xcee

denc

e ra

te, n

Page 67: Ce 490 Lecture Notes-combined

67

In Poisson distribution the probability of x occurrences of an earthquake above a

chosen target magnitude (Mtarget) in t years is

!

(4)

where ν (M>Mtarget) is the annual exceedance rate (or annula frequency) corresponding to

Mtarget in the recurrence model. If no occurrence of events above Mtarget is of interest, x will be

equal to 0 and Eq. (4) will become

0

(5)

On the other hand, in hazard assessment the concern is rather the probability of at least

one earthquake occurring with magnitude above Mtarget. This probability is called as the

exceedance probability (i.e., P(M > Mtarget) and computed by subtracting the probability of

non-occurrence for M > Mtarget (i.e., Eq. (5)) from 1.

1

(6)

By taking the logarithm of both sites in Eq. (6) and writing it for ν (M>Mtarget), one

can obtain the Mtarget for the computed annual exceedence rate. The modified form of

Equation (6) for the computation of annual rate of exceedance is given below.

(7)

In essence, for a time span (exposure time) of t = 50 years and for 10 percent

probability of exceedance of events with magnitudes greater than Mtarget (i.e., P (M>Mtarget) =

0.1), the annual exceedance rate is approximately 2.1e-3 (computed from Eq. (7)). If the

computed annual exceedance rate is used as input in the recurrence model presented in Figure

4.3, the corresponding target magnitude is approximately Mtarget = 6.2 (see blue dashed

arrows in Fig. 4.3). This simple calculation indicates that within a 50-year time period from

today, the likelihood of having an event with M > 6.2 is 10% for the area considered in

Page 68: Ce 490 Lecture Notes-combined

68

seismic hazard analysis. The validity of this conclusion is fairly definitive unless the

applicability of Poisson process is questioned for the area of interest.

4.3. GROUND MOTION PREDICTION EQUATIONS

Ground-motion prediction equations (GMPEs) are derived for estimating the ground-

motion intensity parameters (e.g., PGA, PGV, spectral ordinates Sa at different vibration

periods etc.) to define the level of ground motion in the area of interest. GMPEs describe the

variation of ground motion in terms of important seismological parameters such as magnitude

(M), source-to-site distance (R), site class (SC) and style-of-faulting (SoF). The general

format of GMPEs is given in Eq. (8).

(8)

The functions f(M), f(R), f(SC) and f(SoF) describe the influence of these independent

parameters on the ground-motion intensity parameter. The distribution of ground-motion

intensity parameter (Y) is assumed to be log-normal and the above functional form is

computed through complicated regression analysis by considering the variation of intensity

parameter on the chosen seismological parameters (i.e., M, R, SC and SoF). The uncertainty

about the fitted curve is accounted for by the standard deviation (σ), and ε defines the number

of standard deviations above or below this curve. In this respect GMPEs are probability

distributions conditioned on the seismological parameters used in their derivation. Given a set

of predetermined values of seismological parameters (e.g., M = 6, R = 10 km, Rock site and

strike-slip faulting), GMPEs provide information about the exceedence or non-exceedence

probability of the estimated ground-motion parameter [i.e., P(Y ≤ y|M,R,SC,SoF)].

(9)

Eq. (9) can also be described by the probability distribution of ε

(10)

where fε(ε) is the standard normal distribution with mean zero and variance 1. The theoretical

framework given in Eq. (10) is presented schematically in Fig. 4.5. Under the light of these

explanations, when ε takes a value of 0 the estimated ground-motion parameter has 50%

εσ++++= )()()()()log( SoFfSCfRfMfY

∫∞

−=≤y

YSoFSCRMdYSoFSCRMYfyYP ),,,,(1)(

,,,

∫∞

−=≤*

)(1)(,,,

εε εε dfyYP

SoFSCRM

Page 69: Ce 490 Lecture Notes-combined

69

probability of non-exceedance (i.e., P(Y ≤ y|M,R,SC,SoF) = 0.5). That is why the fitted curve with

ε = 0 is called as the “median curve.” When ε = 1 [i.e., median + σ estimation from Eq. (8)],

P(Y ≤ y|M,R,SC,SoF) is approximately 84%.

Fig. 4.5. Standard normal probability distribution of ε to describe the non-exceedance probability calculations in GMPEs.

Fig. 4.6 describes this concept further by directly using the variation of a ground-

motion prediction equation derived for estimating the PGA (peak ground acceleration). The

regression model is fitted on the rock-site PGA values that are observed from several

recordings with magnitudes ranging from 6.0 and 7.0. The solid curve describes the median

variation of the estimated PGA as a function of distance. The dashed curves show the PGA

estimations for median ± σ. The exceedance probability of the estimated PGAs is ~16% for

the median + σ curve, whereas this probability is ~84% for the median - σ curve. The

probability plot associated with the PGA curves illustrates these interpretations better. Note

that the logarithm of a log-normally distributed variety is normal. Since the ordinate axis in

Fig. 4.6 is given in the log-scale, the presented probability distribution turns out to be a

normal distribution. Fig. 4.7 shows the median and median ± σ pseudo-spectral acceleration

(PSa) ordinates computed from another ground-motion prediction equation for a stiff site

located 10 km away from the causative seismic source (i.e., fault). The scenario event is

considered to be a strike-slip earthquake with M = 6.5.

ε4 -3 -2 -1 0 1 2 3 4

fε(ε)

ε∗

P(ε > ε*)

1 – P(ε > ε*) = P(Y ≤ | )

Page 70: Ce 490 Lecture Notes-combined

70

Fig. 4.6. Median and median ± sigma curves of PGA for M = 6.5 and their comparisons with

the observed stiff site PGA values of 6.0 ≤ M ≤ 7.0.

Fig. 4.7. Median and median ± sigma pseudo-acceleration spectral ordinates estimated by a ground motion predictive model. The seismological parameters used as inputs in the predictive model are

given in the upper right side. Spectral ordinate values at T = 0 correspond to PGA at different sigma levels.

4.4. DETERMINISTIC SEISMIC HAZARD ANALYSIS (DSHA) The main objective of DSHA is to consider a controlling scenario for defining the

hazard in the site of interest. The ground-motion intensity parameter of interest (i.e., PGA, Sa

1

10

100

1000

1 10 100

Pea

k G

roun

d A

ccel

erat

ion,

PG

A (c

m/s

2 )

Distance (km)

median - σ

median

median + σ

0

200

400

600

800

1000

1200

0 0.5 1 1.5 2 2.5 3

Pse

udo-

spec

tral a

ccel

erat

ion,

PS

A(c

m/s

2 )

Period (s)

M = 6.5, R = 10 kmStiff Site, Strike-slipmedian + σ

median

median - σ

Page 71: Ce 490 Lecture Notes-combined

71

at a given vibration period, etc) computed from the controlling scenario corresponds to the

design ground motion. DSHA describes the controlling event for a predetermined magnitude

and location (i.e., source-to-site distance M and R) and specifies the ground-motion

probability level either as median or median + σ estimation by using a suitable ground-motion

prediction equation. The chosen ground-motion prediction equation must be appropriate for

the seismotectonic features of the site where SHA is going be implemented. Figure 4.8 shows

a schematic illustration to explain the major steps in DSHA. The bullets following this figure

describe these steps.

Fig. 4.8. An illustrative figure that shows the basic steps followed in DSHA.

• Define the seismic sources for the area of interest (i.e., S1 and S2 are active faults

likely to affect the hazard in the site). If possible, identify the faulting mechanisms of

these sources as well.

• Estimate the magnitudes of maximum probable events that can occur on the identified

seismic sources (e.g., (Mmax)1 and (Mmax)2). This information can be obtained from the

compiled past and contemporary catalogs that list the temporal and spatial distribution

of events associated with the sources. The alternative can be the use of empirical

magnitude (M) vs. fault rupture length (lrup) relationships (e.g., log M = a+b× lrup

equation developed by Wells and Coppersmith, 1994). The latter option assumes the

entire fault length (or a significant portion of it) ruptures during the controlling event.

In some cases the analyst adds 0.5 magnitude unit on the estimated maximum

magnitude (or a fraction of σ from the empirical magnitude vs. rupture length

relationship) to account for the uncertainty in the magnitude.

• Determine the shortest source-to-site distance between the identified sources and the

site. These are represented as R1 and R2 for S1 and S2, respectively in Figure 4.8a.

(Note: A precise source-to-site distance calculation requires the fault geometry. Such

information is not always available from geological studies. The analyst should make

some proper and physically justifiable assumptions in the absence of such information.

Site

S1,(Mmax)

S2,(Mmax)

R1

R2

Area of interest

S1,(Mmax)1, R1

S2,(Mmax)2, R2

T*

PSa1

PSa2

Period (s)

PSa(a) (b)

Page 72: Ce 490 Lecture Notes-combined

72

The existing methodologies for such kind of analysis are not covered in this chapter

since they are not within the scope of these notes).

• Determine the soil conditions in the site through in-situ geotechnical studies (e.g.,

borehole logs). The soil conditions can be classified in terms of generic site classes

such as rock, stiff or soft soil that are compatible with the code provisions. The soil

conditions can also be determined in terms of the average shear-wave velocity in the

upper 30 m soil profile (VS30) computed through in-situ geophysical analysis. Many

seismic design codes classify site classes by using different intervals of VS30 values.

• Use a proper ground-motion prediction equation to estimate the ground-motion

parameter of interest using the magnitude, distance, site class and style-of-faulting

information obtained in the previous steps. As it is the case in Fig. 4.8.b, the interested

ground-motion parameter can be the pseudo-spectral acceleration (PSa) ordinate at the

fundamental period (T*) of the structure to be designed or to be assessed for its seismic

performance against the controlling event. In general the ground-motion parameter of

interest is either computed as the median (i.e., ε = 0) or median + σ (i.e., ε = 1)

probability level using the chosen GMPE. Some analysts prefer using the median

values of ground motions whereas others chose the latter probability level.

• Compare the ground-motion parameters (PSa(T*) in the given example) computed

from each source and chose the largest one to be used as the design ground-motion

parameter. The corresponding scenario is the controlling event specific to the

considered design or performance assessment project. Note that if the median ground

motion is chosen, the ground motion has 50% chance of exceeding provided that the

controlling event occurs. In a similar manner, the median + σ ground motion has 16%

chance of being exceeded if the controlling event hits the site from the calculated

shortest source-to-site distance.

4.5 PROBABILISTIC SEISMIC HAZARD ANALYSIS (PSHA)

The probabilistic seismic hazard approach accounts for all possible earthquake

scenarios and ground-motion probability levels that can occur on the seismic source(s) in the

area of interest. The probability levels are determined as a range of sigma values above or

below the median ground-motion level. In a sense PSHA can be considered as a large number

of deterministic scenarios (i.e., all possible magnitude, distance and sigma combinations on

Page 73: Ce 490 Lecture Notes-combined

73

the source) that aims at computing the frequency of the ground-motion intensity parameter of

interest (i.e., PGA, PSA at a given vibration period etc) exceeding a specific threshold level

during the chosen exposure time of the structure.

The simplest probability-based computation of annual frequency (γ) for ground-

motion intensity parameter Y, exceeding a threshold level y, from source j can be described by

the following integral:

(11)

In Eq. (11), (νM,min)j is the annual rate of earthquakes with magnitude greater than Mmin for

source j. The probability density functions of magnitude recurrence and distance are described

by fM,j(M)and fR,j(R) for source j, respectively. The concept of magnitude recurrence

relationships for seismic sources is discussed in Section 4.2. The probability density function

fR,j(R) considers the location uncertainty of events occurring on source j and it is generally

assumed to be uniformly distributed. P(Y>y|M,R,SC,SoF) accounts for the variability in the

ground-motion intensity parameter of interest. It is the probability of ground-motion

parameter Y exceeding a threshold level y conditioned on the given magnitude (M), distance

(R), site class (SC) and faulting mechanism (SoF). This probability is determined from the

chosen ground-motion predictive model (see discussions in Section 4.3). Alternatively,

P(Y>y|M,R,SC,SoF) can be written in terms of ε (Eq. (10)) to define the uncertainty in ground-

motion estimations explicitly. Therefore one can express the above hazard integral as

In Eq. (12), P(Y>y|M,R,SC,SoF,ε) is the probability that the ground-motion parameter exceeds the

threshold level y for a predetermined values of M, R, SC, SoF and the number of standard

deviations, ε. This term either attains a value of 0 or 1 depending on the value of estimated

ground-motion parameter exceeding the threshold value y. In other words, the probability

term sorts the scenarios into those that produce ground motions greater than y and those that

produce ground motions smaller than y.

The approach held in Eq. (12) is similar to the methodology followed in DSHA. In

DSHA the magnitude, distance and ε (either taken as 0 or 1) are specified for the controlling

event. That is, there is only one M-R-ε combination. Eq. (12) accounts for all possible M-R-ε

combinations that are likely to influence the site (i.e., many deterministic scenarios). It

( ) ∫ ∫∞

>=>j

j

jRSoFSCRMjR

M

MjMjMj dMdRyYPRfMfyY

min,

max,

min,

)( )( )()(,,,,,min,νγ

( ) )()(f )( )()(min,

max,

min,

max

min

,,,,,,min, ∫ ∫ ∫∞

>=>j

j

jR

M

MSoFSCRMjRjMjMj ddMdRyYPRfMfyY

ε

εεε εενγ (12)

Page 74: Ce 490 Lecture Notes-combined

74

calculates and ranks their rates by considering the temporal, spatial and ground-motion

variability through specific probability distributions. As depicted from Eq. (12), given the

source j, the rate of each M-R-ε combination is

(13)

The integrals in Eq. (12) sum up the rates of the scenarios that produce Y > y. If the

considered site is under the influence of multiple seismic sources the total annual frequency

for Y > y (i.e., γ (Y > y)) is computed by considering the contributions from all of these

seismic sources.

(14)

Implementation of the above theory is explained in the following steps by the help of

representative illustrations given in Fig. 4.9. The major steps in PSHA follow the figure.

Fig. 4.9. Main components of PSHA.

• Develop a set of relevant scenario earthquakes by considering the seismic activity of

the sources in the considered area. This step requires the determination of minimum

and maximum magnitude events that is discussed in Section 4.2. The analyst considers

a range of discrete scenario magnitudes between Mmin and Mmax that are likely to occur

on the source.

• The specific magnitude-recurrence relationship derived for the source (Fig. 4.9.b) will

yield the annual exceedance rate of each scenario magnitude between Mmin and Mmax.

• For each scenario magnitude estimate the rupture length (lrup) from the empirical M vs.

lrup relationships. Depending on the size of rupture, the considered scenario earthquake

can occur at different locations along the source. As illustrated in Fig. 4.9.a n different

εενε ε dfdRRfdMMfRMrate jRjMjM )( )( )()(),,( ,,min,=

∑ >=>j

j yYyY )()( γγ

Gro

und

Mot

ion

(log-

scal

e)

Distance, R (log-scale)

rup1 rup2 rup3 rup4 rupn …

R1 R4 Rn …

SitArea of interest

P(location|M,scenario) = 1/n P(location|M,scenario)

location

M Mscenario

f(M)

P(Y>y|M,R,SC,SoF)

(a) (b) (c)

Page 75: Ce 490 Lecture Notes-combined

75

discrete rupture locations can be set for a given scenario magnitude. Assuming that the

rupture location is uniformly distributed, probability of locations conditioned on the

scenario magnitude (Mscenario) is 1/n [i.e., P (location|Mscenario) = 1/n].

• Compute the shortest distance between each scenario magnitude location and the site

as shown in Fig. 4.9.a. (Refer to the discussions in Section 4.4 about the computation

of source-to-site distances).

• For each scenario magnitude and for the corresponding shortest distances that are

computed from the set of discrete rupture locations, calculate the ground-motion

parameter of interest at the site using a proper GMPE by considering a range of ε

values (Figure 4.9c). The chosen ε range accounts for the inherent variability in

ground motions. In general the ε range is considered ± 3 in PSHA. This range covers

almost 99% uncertainty in the estimated ground motions since ε has a standard normal

distribution. (Refer to any statistical text book to find P(-3 ≤ ε < 3) ≈ 0.99). Note that

the chosen GMPE should account for the seismotectonic features of the area under

consideration and it should properly reflect the site classification of the site. (The

characterization of site class is discussed in Section 4.4).

• Compute the rate for each M, R and ε using the discrete form of Eq. (13).

Note that each rate and the corresponding estimation of ground-motion intensity

parameter of interest form a pair. Considering the number of scenario magnitudes, the

discrete rupture locations for each scenario magnitude and the epsilon range, one can

easily realize the large number of deterministic analysis at the site of interest.

• Finalize the integration scheme in Eq. (12) by ranking the above pairs from the most

severe to the least severe event (described as events associated with the largest and

smallest values of the ground-motion intensity parameter of interest, respectively).

Sum the ranked rates and plot them against the ground-motion levels to find the rates

at which the ground-motion intensity parameter is exceeded for a threshold level. This

plot is called as the “hazard curve”. An illustrative example is given in Fig. 4.10 for

PGA as the ground-motion intensity parameter.

)( )( )()(),,( ,min,, lkscenarioRscenarioMMlkscenarioscenario PRPMPRMrateiiii

ενε ε=

Page 76: Ce 490 Lecture Notes-combined

76

Fig. 4.10. A sample hazard curve computed for PGA to calculate the annual frequency for a given

threshold level.

As discussed throughout this section, PSHA is a probabilistic process that gives

information about the rate of each possible event on a seismic source. The summed rates

constitute the hazard and provide alternatives to describe the annual frequency of ground

motion levels exceeding a certain threshold. For example, the hazard curve presented in Fig.

4.10 depicts that the annual rate of PGA exceeding 0.15g is 0.01. In other words the average

time between the occurrences of this PGA level is 100 years (1/0.01) for the site at which the

specific hazard curve (Fig. 4.10) is calculated. The average time (reciprocal of annual rates

given in the hazard curve) between the occurrences of a certain ground-motion intensity level

is called “return period” in PSHA. A common consensus in the engineering community is to

use ground-motion intensities corresponding to a return period of 2475 years (i.e., γ = 1/2475

= 0.000404) for maximum considered earthquake (MCE) in seismic prone regions. Another

common return period used in earthquake resistant design is 475 years (i.e., γ = 0.0021) to

describe the ground-motion intensities for maximum design earthquake (MDE). MCE and

MDE level ground-motion intensities are used for the design and seismic performance

assessment of different structural systems that are described in the seismic provisions. If the

occurrence of ground-motion intensities is also assumed to be Poisson as in the case of

earthquake occurrence, the return periods for MDE and MCE ground-motion intensities can

be expressed in terms of different probabilities for an exposure time of t years. Using

1.0E-08

1.0E-07

1.0E-06

1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E+00

1.0E+01

1.0E-04 1.0E-03 1.0E-02 1.0E-01 1.0E+00

Ann

ual r

ate

of e

xcee

danc

e

Ground-motion parameter of interest (PGA in g)

Page 77: Ce 490 Lecture Notes-combined

77

Equation (7) and assuming t = 50 years, the exceedance probability for MCE ground-motion

intensities is calculated as 2%. That is, the MCE level ground-motion intensities have 2%

exceedance probability for a service life of 50 years. The same exceedance probability is 10%

for MDE level ground-motion intensities.

4.6 UNIFORM HAZARD SPECTRUM

The last topic of this chapter is to introduce the uniform hazard spectrum (UHS)

concept that is used to define various levels of design ground motion in the current seismic

design codes. Hazard curves derived for a set of pseudo-acceleration spectral ordinates are

used in the construction of UHS. The spectral ordinates corresponding to a predefined annual

frequency (e.g., γ = 0.000404 for MCE level ground motion or γ = 0.0021 for MDE level

ground motions) are identified from the hazard curves computed for a band of vibration

periods. These spectral values constitute the ordinates of the uniform hazard spectrum (Fig.

4.11) and they pose the same annual rate of exceedance (or annual frequency) for the covered

period band.

Fig. 4.11 Uniform hazard spectrum concept: the spectral ordinates of the spectrum given in the lower

row describe the same level of annual frequency (i.e., annual rate of exceedance).

…..

PGA

γ

γMCE

PGAMCE

PGAMCE

PSa

γ

γMCE

PSa(T1)MCE

PSa(T1)MCE

PSa

γ

γMCE

PSa(Tn)MCE

PSa(Tn)MCE

T1 TnPeriod (s)

PSa

Page 78: Ce 490 Lecture Notes-combined

78

Current seismic codes base their design spectra on the uniform hazard spectrum

concept. They identify few spectral ordinates with annual frequencies corresponding to MCE

or MDE level ground motions. The chosen spectral ordinates control the variation of design

spectrum in the short and long period spectral ranges. For example, the International Building

Code (IBC, 2009) defines spectral ordinates at T = 0.2 s and 1.0 s for MCE level ground

motion to establish the design spectrum in the U.S. The Turkish Earthquake Code (TEC,

2007) provides a normalized spectral shape and scales the spectral ordinates by PGA that is

defined for a return period of 475 years. The 475-year PGA levels in the Turkish seismic code

are given for four different seismic zones that are described in the seismic zones map of

Turkey (Fig. 4.12). The seismic zones in this map are identified by the activity of seismic

sources in Turkey.

Fig. 4.12. Turkish seismic zones map

FURTHER READING Edin, 2000.

Ambraseys and Finkel, 2000

Page 79: Ce 490 Lecture Notes-combined

79

Youngs & Coppersmith (1985)

McGuire (2004)

Bommer JEE paper (2002)

Chapter 2 of TEC (2007)

IBC (2009)

Page 80: Ce 490 Lecture Notes-combined

80

5.

Seismic Analysis Procedures in Earthquake Codes

5.1 INTRODUCTION

Seismic analysis procedures in earthquake codes prescribe calculation of seismic

design forces. These procedures mainly comprise of the following steps:

1. Calculate lateral earthquake forces for linear elastic response.

2. Reduce linear elastic forces to account for inelastic response.

3. Apply reduced forces to the structure, carry out structural analysis and determine the

internal seismic design forces acting on structural members.

4. Combine internal seismic design forces with the internal forces from gravity loads.

5. Design structural members under these combined design forces.

In Step 1, lateral earthquake forces are calculated from a linear elastic acceleration

response spectrum which represents design ground motion intensity. The linear acceleration

spectrum is reduced in Step 2 in order to account for the inelastic deformation capacity

(ductility) of the system. Then a structural model is constructed in Step 3, and response

spectrum analysis is carried out under the reduced acceleration spectrum (design spectrum).

The internal seismic design forces determined in Step 3 are combined with the results of

gravity analysis in Step 4. Finally structural members are designed under these combined

force demands in Step 5. Hence, this is a “force-based” design procedure since the design of

structural members is based on internal forces which indirectly account for the inelastic

deformation capacity of the conceived structural system. This deformation capacity has to be

“assumed” in advance since the system has not been designed yet.

Although the analytical development of the procedures introduced in this Chapter is

similar to the procedures discussed in Chapters 2 and 3, their format is presented in

conformance with the general code notation.

Page 81: Ce 490 Lecture Notes-combined

81

5.2 DESIGN GROUND MOTIONS: LINEAR ELASTIC RESPONSE SPECTRUM The design earthquake is expressed by a smooth linear elastic acceleration response

spectrum in seismic design codes. A smooth spectrum represents a statistical average of the

response spectra obtained for many ground motions with similar intensities, recorded on

similar sites. We may consider the mean spectrum of ground motions normalized with respect

to PGA in Fig. 2.18.b which are recorded on semi-stiff soils. This figure is repeated below in

Fig. 5.1 for convenience.

Fig. 5.1. Mean acceleration spectra of 10 ground motions normalized with respect to their peak ground accelerations.

5.2.1 Linear Elastic Acceleration Spectrum in the Turkish Code The normalized mean spectral curve in Fig. 5.1 can be approximated by a simple

smooth curve for design purposes. The smooth normalized mean spectral shape is expressed

by a generic spectrum shape in the 2007 Turkish Earthquake Design Code, which is

composed of three line segments. The generic normalized spectrum shape is shown in Fig.

5.2.

Fig. 5.2. Normalized spectrum shape in the Turkish Seismic Code, 2007.

Page 82: Ce 490 Lecture Notes-combined

82

The three portions of the linear elastic acceleration spectrum in Fig. 5.2 are expressed by

1 1.5

2.5 (5.1)

2.5 .

Soil conditions at the site are represented by TA and TB in the Turkish Earthquake Code

which are defined in Table 5.1. As the soil conditions vary from very stiff to soft, it is

reflected in the spectrum shape by a longer plateau region. This approximation for reflecting

the effect of soil conditions on spectral shape is under strong criticism lately since it is known

that the amplification factor 2.5 in Eq. (5.1) is also sensitive to the local soil conditions.

Table 5.1. Corner periods for different soil conditions

Local Soil Conditions TA (s) TB (s) Z1 (rock) 0.10 0.30 Z2 (stiff soil) 0.15 0.40 Z3 (soft soil) 0.15 0.60 Z4 (loose alluvium) 0.20 0.90

A comparison of the normalized spectrum of the Turkish Earthquake Code for Z2 soil

type with the normalized spectrum in Fig. 5.1 is shown in Fig. 5.3. The agreement is evident

except at very small periods in the 0.10 second range (stiff structures with 1 or 2 stories)

where the Code spectrum is more conservative.

Fig. 5.3. Normalized linear elastic acceleration spectrum in the Turkish Earthquake Code (2007) for stiff soil conditions (Z2), compared with the normalized mean linear elastic acceleration spectra from

10 ground excitations.

Page 83: Ce 490 Lecture Notes-combined

83

When the normalized spectrum is scaled by the peak ground acceleration PGA along the

vertical axis and the structure importance factor I is applied, an absolute spectrum is obtained.

The absolute spectrum is expressed by Eq. 5.2 and plotted in Fig. 5.4 where PGA= , is

the acceleration of gravity (9.81 m/s2), is given in Eq. (5.1) and TA and TB are the corner

periods defined in Table 5.1.

· · (5.2)

Fig. 5.4. Absolute linear elastic acceleration spectrum in the Turkish Seismic Code, 2007 (I=1).

The coefficient in Eq. 5.2 and in Fig. 5.4 is called the “effective ground

acceleration coefficient”. It is determined by the probabilistic seismic hazard analysis (PSHA)

procedure discussed in Chapter 4. is given in Table 5.2 for five different seismic zones in

Turkey. Five seismic zones in Table 5.2 were defined in the “Seismic Zone Map of Turkey”

shown in Fig. 4.12.

Table 5.2. Effective ground acceleration coefficient in the five seismic zones in Turkey

Seismic Zone 1 2 3 4 5

0.40 0.30 0,20 0.10 0.0

5.2.2 Structure Importance Factor (I)

The structure importance factor I in the above equation indirectly account for less risk,

or better expected performance specified for important structures. It increases the seismic

design forces by 1.5 for emergency facilities (hospitals, fire and police stations, etc.), by 1.4

for schools and by 1.2 for stadiums, theatres and concert halls, compared to the ordinary

buildings. Accordingly these structures are designed for higher lateral strength, and hence

Page 84: Ce 490 Lecture Notes-combined

84

they are expected to sustain less damage under the design earthquake. Emergency facilities (I

=1.5) are expected to remain functional after a severe earthquake.

5.3 REDUCTION OF ELASTIC FORCES: DESIGN SPECTRUM We have discussed earlier in Chapter 2 that linear elastic spectrum can be reduced if

inelastic response is permitted. The reduced inelastic acceleration response spectrum Sai  (T)

can be directly obtained from the linear elastic acceleration response spectrum Sae (T) through

μ (5.3)

by invoking Eq.(2.65.a). Rμ  is the ductility reduction factor in the above equation, which was

introduced in Section 2.6.4. It is a function of both period and the ductility factor (see Figures

2.28 and 2.29).

This force reduction concept for SDOF systems introduced in Chapter 2 can also be

applied to the MDOF systems which can be idealized as equivalent SDOF systems. Let’s

consider a frame in Fig. 5.5 with inelastic deformation capacity, subjected to an increasing

lateral force distribution (static pushover analysis).

(a) (b) Fig. 5.5. An inelastic frame subjected to increasing static lateral forces (pushover analysis).

a) Incremental lateral forces and corresponding incremental lateral displacements, b) idealized inelastic force-deformation (moment-curvature) characteristics of the yielding member ends.

For each force distribution , there is a corresponding displacement distribution ui

obtained from nonlinear static analysis. The distribution of can be taken similar to the

distribution of equivalent static first mode force distribution given by Eq. (3.50) with n=1, but

Page 85: Ce 490 Lecture Notes-combined

85

applied incrementally in small lateral load steps. The sum of lateral forces in is the base

shear force Vbi where 1 · , and the roof displacement at top is uroof,i  at the i’th load

step.

When the loading increments are applied progressively on the inelastic structure,

internal moments at certain sections will eventually reach their yield capacities shown in Fig.

5.5.b with black dots, and these sections will enter the elasto-plastic response region. As

further lateral load increments are applied, such sections will not be able to develop additional

resistance and they redistribute the additional internal forces that they cannot resist to other

sections while exhibiting plastic deformation. Hence, the number of plastic end regions

(plastic hinges) increase with increasing lateral loads, accompanied by reduction of lateral

stiffness of the system. The state of plastic hinge distribution at three different lateral load

states is depicted in Fig. 5.6.

Fig. 5.6. Plastic hinge distribution in an inelastic frame subjected to increasing static lateral forces

(pushover analysis) at three different loading states.

If we plot Vb versus uroof, we obtain a capacity curve which resembles the force-

displacement relation of an equivalent SDOF inelastic system. Capacity curve for the frame in

Fig. 5.6 is shown in Fig. 5.7, and the lateral load states indicated in Fig. 5.6 are also marked

on the capacity curve.

Various strength levels are indicated on the capacity curve in Fig. 5.7. Ve is the base

shear force demand from linear elastic structure by the earthquake ground motion which is

expressed by a linear elastic spectrum. Vy is the yield base shear force or the base shear

capacity whereas Vd is the presumed design base shear of the frame structure.

Page 86: Ce 490 Lecture Notes-combined

86

Fig. 5.7. Capacity curve of the inelastic frame obtained from pushover analysis.

We may now introduce two ratios among these strength levels.

μ : Ductility reduction factor (5.4)

: Overstrength reduction factor (5.5)

Ductility reduction factor Rμ in Eq. (5.4) that is defined for an inelastic MDOF system

is identical to the ductility reduction factor introduced above in Eq. (5.3), and previously

introduced in Chapter 2 for an inelastic SDOF system. The similarity between Fig. 5.7 for an

inelastic MDOF system and Fig. 2.27 for an ideal inelastic (elasto-plastic) SDOF system is

also evident.

The actual yield base shear force (yield strength) Vy of a system that is designed for a

design base shear force of Vd is usually larger than Vd (i. e. ) because of the

overstrength inherent in design. The overstrength reduction factor Rov describes this deviation

from the target strength. Various factors contributing to overstrength are;

• material strength reduction factors for concrete, etc. ,

• minimum dimensions,

• minimum reinforcement ratio in RC members,

• minimum strength of materials,

• detailing requirements,

• redundancy (redistribution of internal forces from the yielded to non-yielded

members in indeterminate systems).

Page 87: Ce 490 Lecture Notes-combined

87

We can simply combine Rμ and Rov in Eqs. (5.4) and (5.5) into a single reduction factor.

μ x . (5.6)

Ra (T) is called the earthquake force reduction factor. In the 2007 Turkish Earthquake Code,

Ra (T) is defined by a bi-linear spectral curve shown in Fig. 5.8.

Accordingly, the reduced design spectrum SaR is defined by

(5.7)

which is similar to Eq. (5.3).

Fig. 5.8. Earthquake force reduction factor in the Turkish Earthquake Code (2007).

Ra (T) in the Turkish Earthquake Code depends on another factor R as shown in Fig.

5.8, which is called the structural system response factor. When Fig. 5.8 is compared with

Fig. 2.29, we may observe that R represents the ductility capacity for an equivalent inelastic

SDOF system, and TA is the corner period of the design earthquake ground motion. The R

factors in the Code include both ductility and overstrength however as shown in Eq. (5.6).

R factors are described in Table 2.5 of the 2007 Turkish Earthquake Code for various

structural systems and for two ductility levels, one with ordinary ductility and the other for

enhanced ductility implemented in design. These ductility levels are imposed by capacity

design requirements discussed in the following Chapter. The R factors in Table 2.5 are based

on judgment as well as experimental and analytical verifications.

The reduced design spectra in the Turkish Earthquake Code for R=4 (ordinary RC

frame) and R=8 (ductile RC frame) are shown in Fig. 5.9, together with the linear elastic

design spectrum.

Page 88: Ce 490 Lecture Notes-combined

88

Fig. 5.9. Earthquake design spectra in the Turkish Earthquake Code (2007) for linear elastic (R=1), ordinary (R=4) and ductile (R=8) RC frames.

5.4 ANALYSIS PROCEDURES

There are two basic linear elastic analysis procedures in seismic design codes which

are carried out under the design earthquake represented by a reduced design spectrum SaR(T).

The first one is the modal superposition procedure and the second one is the equivalent static

lateral load procedure, which is derived from the first one as a special case. It is applicable to

simple, regular structures.

In both seismic analysis procedures, only those degrees of freedom which inertial

forces act are considered in the equations of motion. These are called the dynamic degrees of

freedom. The remaining static degrees of freedom are necessary for calculating the internal

forces under gravity loads. However they can be eliminated from the dynamic equations of

motion by a method called static condensation, which leads to a significantly reduced number

of degrees of freedom compared to the original system.

5.4.1 Reduction of Degrees of Freedom: Static Condensation

When the slabs act as rigid diaphragms in their own plane, then the motion of each

slab during an earthquake can be defined by three dynamic degrees of freedom defined at its

centre of mass. On the other hand, almost the entire mass of a building is confined to the slabs

(concrete slab, cover, beams, live loads, hung ceilings, etc). Only columns and partition walls

are not a part of the slab system; however their masses can be distributed evenly to the

Page 89: Ce 490 Lecture Notes-combined

89

adjacent slabs at top and bottom. Accordingly, the inertial forces which act on the story

masses during an earthquake can be defined on the slabs.

Let’s consider a rigid slab with mass m moving in its own plane. Two consecutive

positions of the rigid slab are shown in Fig. 5.10.a with the local coordinate axes x-y and x’-

y’, respectively. The motion of the centre of mass in the plane of motion is sketched in Fig.

5.10.b. The motion of the corner point A (uxA  ,  uxB) can be expressed in terms of the

translational motion of the centre of mass (ux , uy), rotation of the rigid slab about its centre

of mass, and the position of A with respect to the centre of mass (xA , yA).

· (5.8.a)

· (5.8.b)

(a) (b)

Fig. 5.10. Motion of a rigid slab in its own plane.

Hence, the inertial forces acting on a rigid floor slab can be expressed in terms of two

inertial forces fx , fy and a torsional moment Mz .

·

; ·

; ·

(5.9)

In Eqs. (5.9),

,

and

Page 90: Ce 490 Lecture Notes-combined

90

The total and relative rotations are equal since the ground motions do not have rotational

components generally. for rectangular slabs with dimensions a and b.

ux , uy and θz at each floor are considered as the dynamic degrees of freedom (3xN for

an N-story 3D frame) whereas all other DOF’s are called the static degrees of freedom

because a mass is not assigned to them (column and beam end rotations, slab edge rotations,

axial deformations). Dynamic forces do not act on static DOF’s.

We can eliminate the static degrees of freedom during dynamic analysis by a

procedure called static condensation. Let’s define the displacements along dynamic DOF’s by

ud and those along static DOF’s by us. The dynamic and static DOF’s of a single story, multi-

bay frame are sketched in Fig. 5.11.

Fig. 5.11. Dynamic and static DOF’s of a single story, multi-bay frame.

Accordingly, the vector of inertial forces acting on dynamic DOF’s is whereas =

0 since no dynamic forces act on the static DOF’s. Then the stiffness equation between forces

and displacements under dynamic excitation becomes,

0 (5.10)

The second row of Eq. (5.10) can be expanded as

0 (5.11)

which yields

(5.12)

Substituting from Eq. (5.12) into the first row of Eq. (5.10) gives

Page 91: Ce 490 Lecture Notes-combined

91

· · (5.13)

The term in the parenthesis in Eq. (5.13) is called the condensed stiffness matrix kd which has

a size of (3Nx3N) in an N-story 3D space frame and a size of N in a 2D plane frame. Static

condensation reduces the size of the matrix equation of motion enormously. The 2D frame in

Fig. 5.11 has a total of 9 DOF’s whereas it reduces to a 1 DOF system after static

condensation.

5.4.2 Mode Superposition Procedure A modal superposition analysis is carried out for the condensed system under the

reduced design earthquake SaR by using a sufficient number of modes. This method is

applicable to all buildings without any restrictions.

Eigenvalue analysis is conducted on the condensed system with stiffness kd in Eq.

(5.13) and with a lumped- mass diagonal matrix m defined in Eq. (5.14). The components of

the mass matrix in a 3D building are assigned at each story, which are the story mass in two

translational directions x and y, and the mass moment of inertia in the rotational direction θz  

along the dynamic DOF’s of the associated slab.

. . .

(5.14)

Hence, mi is the mass of the i’th story, and is the mass moment of inertia of

the i’th slab in Eq. (5.14).

The equation of motion for the condensed system under x or y direction of ground

motion is obtained from Eq. (3.3).

1 or (5.15)

When the right hand side is zero during free vibration, we obtain the eigenvalue problem.

0 (5.16)

Page 92: Ce 490 Lecture Notes-combined

92

The solution of eigenvalue problem gives the eigenvalues and the eigenvectors

of the condensed system. The eigenvector contains two translational components

φ and φ , and one rotational component φ at each floor i of a 3D building. The modal

forces also include two horizontal forces and a torque at each floor accordingly.

φφφ

φφφ

.

.

.

φφφ

...

(5.17)

In a 2D plane frame, the φ and φ components in do not exist.

The modal forces under the reduced design earthquake spectrum follows from Eq. (3.51).

φ , (5.18)

Then we apply in Eq. (5.18) to the “uncondensed” building structure and calculate member

forces and displacements at the n’th mode from static analysis. Two horizontal force

resultants and one torsion moment applied at the centre of mass of the j‘th floor slab of a 3D

structure is shown in Fig. 5.12.

Fig. 5.12. Dynamic force components at the n’th mode, acting at the mass centre of the j’th floor slab.

Page 93: Ce 490 Lecture Notes-combined

93

If the 3D building is symmetrical about the x axis, then = 0 and = 0 under

excitation SaR,x in the x direction for all modes.

Finally, we combine the modal results by SRSS or CQC formulation. For example, the

design moment at the top end of the front right corner column in Fig. 5.12 is calculated from

(5.19)

The Minimum Number of Modes

The total effective mass of the Nmin modes considered in the mode superposition

analysis should be larger than 90% of the total mass, separately in the x and y directions. This

requirement is verified from the following inequalities.

: 0.90 (5.20)

: 0.90 (5.21)

In Eqs. (5.20) and (5.21), N is the number of stories and mi is the story mass. Also,

φ ; φ (5.22)

and,

∑ φ φ . φ . (5.23)

It should be noticed that Eqs. (5.22) and (5.23) are the expanded scalar forms of Eqs. (3.33)

and (3.30) respectively, where

; ; (5.24)

and 1 0 0 ; 1 0 0 ; ; 0 1 0 ; 0 1 0 ; are the

influence vectors in the x and y directions, respectively. and transmit the motion of the

ground in the x and y directions to the respective translational dynamic DOF’s of the structure

above as rigid body motion components.

Page 94: Ce 490 Lecture Notes-combined

94

Summary

The mode superposition analysis procedure in the Code can be summarized with the

following steps.

1. Prepare a complete structural model of your structure.

2. Condense the static DOF’s (Eq. 5.13).

3. Carry out eigenvalue analysis of the condensed structure (Eq. 5.16) and determine

(or Tn) and for each mode n.

4. Calculate modal spectral accelerations from the design spectrum SaR,n   (Eq. 5.7) for a

proper R factor.

5. Calculate the modal forces , (Eq. 5.18).

6. Apply to the complete (uncondensed) structural model by expanding to the

complete force vector as shown in Fig. 5.12, and determine internal modal

member forces and deformations rn from

· where is the full

(uncondensed) stiffness matrix and includes all DOF’s.

7. Combine the modal results by SRSS:

.

8. Further combine gravity and earthquake analysis results:

Example 5.1. Consider the 2 story cantilever frame in Fig. (a). The stiffness matrix for a frame member is given in Fig. (b)

a) Determine the stiffness matrix of the system. b) Apply static condensation for calculating the stiffness matrix of the reduced 2 DOF

system. Inverse of a 2X2 matrix is given in Fig. (c). c) Calculate the eigenvalues and eigenvectors of the reduced system (in terms of EI, L

and m). d) Calculate the EQ moment diagrams for each mode under the TEC 2007 design

spectrum (zone 1, soil type Z3, R=1 and I=1) in terms of mgL if 116

1

12 6 12 66 4 6 2126

62

126

6 4

u3

u1u2

u4

EIL

(a)

u2 θ2

u1 θ1

L EI

L EI

m

m

(c)

(b)

Page 95: Ce 490 Lecture Notes-combined

95

Solution:

a) Stiffness Matrices Global stiffness of the 1st story column;

12 66 4 ;

Global stiffness of the 2nd story column:

12 6 12 66 4 6 2126

62

126

6 4

Mapping the two matrices on the global DOF’s;

24 0 12 60 8 6 2126

62

126

64

We should separate the dynamic DOF’s (u1, u2) and static DOF’s (θ1, θ2) for static condensation. This requires interchanging second and third rows and columns of ksys respectively, which gives;

24 12 0 612 12 6 606

66

82

24

b) Static Condensation

and .

8 22 4

4 11 2 . Considering that ac 8 1 7 from Fig. (c),

2 11 4 . Then the condensed stiffness matrix is obtained from the above

equation as

24 1212 12

0 66 6 2 1

1 40 66 6 , or

u1 θ1

u2

θ1 u1

θ2

Page 96: Ce 490 Lecture Notes-combined

96

24 1212 12

72 5454 72 96 30

30 12 .

c) Eigenvalue Analysis

0 1 00 1

96 30

30 12 1 00 1 0

0 , or, 967

307

307

127

00 .

det ( ) =0 gives; 0. Simplifying and rearranging,

1152 108 900 0 Let and . Substituting and solving the quadratic equation above,

, √ ; 2.386 and 105.614 . Accordingly, 0.341 and 15.088 .

Then 0.584 ; 3.884

Mode vector 1

93.614 3030 9.614 0

93.614 30 0 Let 1.0 Then . 3.12. Hence Ф 1.00

3.12 . Mode vector 2

79.614 30

30 93.614 0

9.614 30 0

Let 1.0. Then 0.32. Accordingly Ф 1.00

0.32 .

Ф1

3.12

1.0

Ф2

-0.32

1.0

Page 97: Ce 490 Lecture Notes-combined

97

d) EQ Moment Diagrams When 116 / , T1=1.0 s and T2=0.15 s. The spectral accelerations for the two modes are obtained from the Code spectrum given below.

0.665 6.52 / ; 1.0 9.81 / The modal force vectors are expressed as . Here,

1 4.12 ; 0.68 and 10.73 ; 1.102 . Hence, 0.384 and 0.617. The modal forces are then,

0.384 1.003.12 0.665 0.255

0.797 and

0.617 1.000.32 1.0 0.617

0.197 . Since the system is statically determinate, we can calculate the moment distributions directly. Moment diagram for Mode 1 Moment diagram for Mode 2 SRSS:

0.197mg

0.617mg

L

L0.797mgL

1.849 mgL

0.797mg

0.255mg

L

L

0.197 mgL

0.223 mgL

0.82 mgL

1.86 mgL

Page 98: Ce 490 Lecture Notes-combined

98

Example 5.2. A two story, one bay frame with six DOF’s is given in Fig. (a). All frame members have a length of L, moment of inertia of I and the modulus of elasticity is E. A uniform distributed vertical gravity load with intensity q is acting on the beams. Determine the design moments and shear forces in the column AC and the beam CD under gravity and earthquake loads. Design earthquake is defined with the reduced design spectrum in the Turkish Earthquake Code 2007 for seismic zone 1, soil type Z3 and an ordinary ductility level, which is given in Fig. (b).

(a) Frame structure (b) Reduced design spectrum Solution The procedure summarized above will be implemented in the solution. Static Condensation The 6 DOF system in Fig. (a) will be reduced to a 2 DOF dynamic system for eigenvalue and mode superposition analyses. Displacement vector u of the unreduced 6 DOF system and the displacement vector ud of the reduced (condensed) 2 DOF system are given below. Hence, the condensed system DOF’s are the lateral DOF’s along the floor levels where the story masses are defined and accordingly the inertial forces develop.

(1)

The stiffness matrix of the 6 DOF system is

48 24 0 24 24 6

0 6 12

0 62

6 6 6 6 2 0

0 6 2 12 0 26 6

66

20

02

8 2

2 8

(2)

The stiffness matrix above can be partitioned as

L

L

L

E

A

F

B

C D

u4

u1

u2

u6 u5

u3

EI

m

m

Page 99: Ce 490 Lecture Notes-combined

99

where the condensed stiffness matrix kd is expressed as kd = kdd – ksd

T kss-1 kds. Then,

17690 300

300 228

Similarly, the mass matrix for the condensed 2 DOF system is

0

0

Eigenvalue Analysis The equation det (kd – ω2 m ) = 0 should be solved for determining the eigenvalues and eigenvectors. Here, we will assign numerical values to frame properties. Let EI = 660 kN.m2 , L = 2 m and m =10 tons. Then, the eigenvalues are obtained as:

ω12 = 39.00 (rad/s)2 ; ω2

2 = 406.50 (rad/s)2 and T1 = 1.00 s ; T2 = 0.31 s. The mode shapes are determined by solving the the equation 0 . The solution gives;

0.49

1.00

2.03

1.00 (3)

Spectral Accelerations Spectral acceleration values for the modal periods are obtained from the reduced design spectrum of TEC 2007 given in Fig. (b) above (R=4, Zone 1, soil type Z3):

Mode 1: T1 = 1.00 s, SaR,1 = 1.63 m/s2

Mode 2: T2 = 0.31 s, SaR,2 = 2.45 m/s2 Modal Force Vectors

, where ; and 11

When the respective values of , and are substituted into these expressions, we find

L1 = 14.92 tons ; L2 = -10.32 tons; M1 = 12.42 tons ; M2 = 51.29 tons, and

9.63

19.58 kN ;

10.03

4.93 kN (4)

Earthquake Analysis (Mode Superposition Procedure) The modal force vectors of the condensed 2 DOF system are first expressed for the uncondensed (original) 6 DOF system in view of the DOF’s defined in Fig. (a).

Page 100: Ce 490 Lecture Notes-combined

100

9.63

19.580000

kN ;

10.03

4.930000

kN (5)

Then the modal displacement vectors are obtained by solving

· where ksys is

the 6X6 global system stiffness matrix in Eq. (2). Solution for n =1 and n =2 gives,

0.02470.05020.00990.00990.00570.0057

;

0.00250.0012

0.00010.00010.00110.0011

(6)

The units are meters and radians.

Earthquake Forces in Member AC Member forces are calculated from the member equilibrium equation;

, · ,

12 6 12 66 4 6 2126

6 2

12 6 6 4

; 00

Mode 1

, 00

00

0.02470.0099

14.61 17.89 14.61 11.32

kN and kN.m

Mode 2

, 00

00

0.00250.0001

2.55 2.51 2.55 2.58

kN and kN.m

uA 0θA 0

θC u3

uC u1

A

C

Page 101: Ce 490 Lecture Notes-combined

101

Earthquake Moment Diagram AC

Mode 1 Mode 2 SRSS

 MAC,bot = (17.892+2.512)1/2 = 18.06 kN.m; MAC,top = (11.32+2.582)1/2 = 11.61 kN.m Earthquake Forces in Member CD Member forces are calculated from the member equilibrium equation;

, · ,

12 6 12 6 6 4 6 2

12 6

6 2

12 6 6 4

;

0

0

Mode 1

,

0

0

00.0099

00.0099

19.7019.70

19.7019.70

kN and kN.m

Mode 2

, 00

00.0001

00.0001

0.21 0.21

0.21 0.21

kN and kN.m

 

17.89  kN.m 

‐11.32  kN.m  

2.51  kN.m

‐2.58  kN.m

18.06  kN.m

‐11.61  kN.m

 

DC

uC 0 uD 0

θC u3 θD u4

Page 102: Ce 490 Lecture Notes-combined

102

Earthquake Moment Diagram CD

Mode 1 Mode 2

SRSS

 MCD,l = (19.702+0.212)1/2 = 19.71 kN.m ; MCD,r = (19.702+0.212)1/2 = 19.71 kN.m Gravity Load Analysis Let q=50 kN/m. This value is consistent with the story mass of 10 tons, hence 10g kN is distributed uniformly over the story girders of L=2 m. Fixed-end moments at beam ends are calculated from MFEM=qL2/12 = 16.67 kN.m. Accordingly, the gravity force vector is

00

16.67 16.67

16.67 16.67

kN. m

Displacement vector under gravity forces is determined by solving

· . The

global displacement vector and the member end displacements are obtained as,

00

0.0036 0.0036

0.0072 0.0072

; 00

000

0.0036

;

0

0 0

0.0036 0 0.0036

Then the member end forces under gravity loads are determined from

 

19.70  kN.m 

‐19.70  kN.m

0.21  kN.m

‐0.21  kN.m

 

19.70  kN.m 

‐19.70  kN.m

q

L= 2 m. θL θR C D

Page 103: Ce 490 Lecture Notes-combined

103

, · and ,

·

where , 0 no span loading and ,

016.67

016.67

kN. m. Then,

,

3.572.383.574.76

and ,

014.30

014.30

Gravity Moment Diagrams of AC and CD AC CD Combining the Internal Forces: MG MEQ

Member AC: MG + MEQ

-2.38 kN.m

4.76 kN.m

-14.30 kN.m -14.30 kN.m

+ =

-6.85kN.m

15.68 kN.m 18.06 kN.m

-11.61 kN.m

-2.38 kN.m

4.76 kN.m

Page 104: Ce 490 Lecture Notes-combined

104

Member AC: MG - MEQ

Member CD: MG + MEQ Member CD: MG - MEQ It should be noted that the design moments in columns and the design shear forces in columns and beams are calculated differently in the capacity design procedure introduced in the following Chapter 6.

-2.38 kN.m

4.76 kN.m

18.06 kN.m

-11.61 kN.m

- =

19.70 kN.m

-19.70 kN.m

+

=

5.4 kN.m

-34 kN.m

-14.30 kN.m -14.30 kN.m

19.70 kN.m

-19.70 kN.m

-14.30 kN.m -14.30 kN.m

-

=

5.4 kN.m

-34 kN.m

16.37 kN.m

-20.44 kN.m

Page 105: Ce 490 Lecture Notes-combined

105

Example 5.3. The system shown below is in seismic zone 1, soil type 3 and I=1 (Turkish Seismic Code). The slab is a rigid diaphragm in its plane and also out of plane. The mass of the slab is 20 tons, and its center of mass is located as shown. Ignore the mass of columns.

a-) Calculate the elastic modal displacements (center of mass) and modal forces under seismic forces acting in the x-direction only.

b-) Determine the design forces for columns under the same excitation. Assume that the columns share the weight of the slab equally. There is no live load, and the design load combination is DL+EQ. Use R=4 in design. Ec=20e6 kN/m2 , Columns: 0.4m x 0.4m, IG=88.33 ton.m2. Free vibration properties of the system are given below. Note that each modal vector consists of three components (3D system), the first one is translation in x, the second one is translation in y, and the third one is rotation about the centre of mass. Solution 0.1595 0.157 0.092 2.5, 2.5, 1.92

39.4 / 40.0 / 68.3 /

0.222

00.0131

0

0.22360

0.02755

00.1056

Page 106: Ce 490 Lecture Notes-combined

106

Modal Displacements

20 0 00 20 00 0 88.33

tons

. . 1 0 0

4.44 , 0, 0.551

. . 4.44 0.2229.81

39.4 6.23 10 , 0,

. . 0.551 0.027557.53

68.3 2.45 10

. . 4.44 0.01319.81

39.4 3.677 10 , 0,

. . 0.551 0.10567.53

68.3 9.4 10

6.23 10

03.677 10

, 0 , 2.45 10

09.4 10

Modal Forces: . Design Forces in Columns

The superscript 1 in the figure indicates the 1’st floor. Since there is only one floor, we can drop this index.

. .

. 0.4 1 2.5 9.81 9.81 /

. 0.4 1 2.5 9.81 9.81 /

. 0.4 1 1.92 9.81 7.53 /

191.650

50.40 ;

2.290

38.73 .

128000 ⁄

For a fixed-fixed square column;

Page 107: Ce 490 Lecture Notes-combined

107

Mode 1 ( 6.23 10 ; 3.677 10 ) Column 1 Column 2 Column 3 Column 4 Mode 3 ( 2.45 10 ; 9.4 10 ) Column 1 Column 2

2 6.23 10 3.677 10 2 6.9654 3

6.9654 3 8000 55.7232

9.1925 4 8000 7.354

. 2 7.354 2 14.708 . 2 55.7232 2 111.446

2.5 3.677 10 2.5 9.1925 4

3 6.23 10 3.677 10 3 5.1269 3

5.1269 3 8000 41.01 ; 7.354 2 14.708

9.1925 4 8000 7.354 ; 41.01 2 82.02

2.5 9.1925 4

3 5.1269 3 41.01 14.708 2.5 9.1925 4 7.354 82.02

2 6.9654 3 55.7232 14.708

2.5 9.1925 4 7.354 111.446

2 2.45 10 9.4 10 2 1.635 10

1.635 10 8000 1.308 1.88 2 3.76

2.35 10 8000 1.88 1.308 2 2.616

2.5 9.4 10 2.5 2.35 10

3 2.45 10 9.4 10 3 3.065 10 2.5 2.35 10

3.065 10 8000 2.452 1.88 2 3.76

2.35 10 8000 1.88 2.452 2 4.904

Page 108: Ce 490 Lecture Notes-combined

108

Column 3 Column 4

Combining the internal forces with SRSS

Column 1

√55.7232 1.308 55.74 ; √7.354 1.88 7.59

√14.708 3.76 15.18 ; √111.446 2.616 111.47 Column 2

√41.01 2.452 41.08 ; √7.354 1.88 7.59

√14.708 3.76 15.18 ; √82.02 4.904 82.16

Column 3

√44.01 2.452 41.08 ; √7.354 1.88 7.59

√14.708 3.76 15.18 ; √82.02 4.904 82.16

Column 4

√1.308 55.7232 55.74 ; √7.354 1.88 7.59

√14.708 3.76 15.18 ; √111.446 2.616 111.47

Columns 1 and 4 are the most critical. Design shear forces (Reduce by R=4): . 13.9 ; . 1.9 Design moments (Reduce by R=4): . 3.8 ; . 27.9

3 3.065 4 2.452 3.76

2.5 2.35 4 1.88 4.904

2 1.635 4 1.308 3.76

2.5 2.35 4 1.88 2.616

Page 109: Ce 490 Lecture Notes-combined

109

Shear Force Diagram Moment Diagram

Axial load for column ① Slab weight 20 9.81 196.2

Columns share slab weight equally ① 49.05

Design will be conducted under; N 49.05 M 3.8

F 1.9 and

N 49.05 M 27.9

F 13.9

For y direction For x direction 5.4.3 Equivalent Static Lateral Load Procedure

Let’s consider the effective modal forces and the resulting modal displacements at the

first mode of a building in Fig. 5.13, as discussed in Section 3.4.2.

Fig. 5.13. Modal forces and resulting modal displacements at the first mode of a building.

Page 110: Ce 490 Lecture Notes-combined

110

The modal force vector under the reduced design spectrum was given in Eq. (5.18).

Then the base shear force at the 1’st mode can be obtained from Eq. (3.53).

, , (5.25) The j’th component of in Eq. (5.18) and Fig. 5.13 can be expressed in scalar form as

φ , (5.26) When we multiply and divide the RHS by and rearrange;

.1

φ , (5.27) Substituting from Eq. (5.25) into Eq. (5.27), we obtain

φ

(5.28)

where 1 ∑ φ

(5.29)

Note that in Eq. (5.29) is identical to its definitions in Eq. (5.22) for n=1 in the x and y

directions, respectively. Finally, substituting from Eq. (5.29) into Eq. (5.28), we get

φ

∑ φ (5.30)

Multiplying and dividing the RHS by the gravitational acceleration g gives,

φ

∑ φ (5.31)

where is the weight of the i’th floor.

When the first mode (in the direction of earthquake excitation) is dominant on total

dynamic response, then

, (5.32)

from Eq. (5.25). Similarly, from Eq. (5.31), considering that ,

Page 111: Ce 490 Lecture Notes-combined

111

φ

∑ φ (5.33)

The components of the first mode vector in the direction of earthquake excitation

can be approximated with a linear variation over the building height in simple buildings with

regular height wise distribution of mass and stiffness.

φ (5.34)

Hj is the height of the j’th floor from the base and is an arbitrary constant representing the

slope of linear distribution. Substituting φ from Eq. (5.34) into Eq. (5.33), we obtain

∑ (5.35)

Base Shear Force in the Code It was proposed in Eq. (5.32) that , in simple regular buildings where

is the effective modal mass of the first mode. If we replace with the total mass

∑ , then we approximately account for the masses of higher modes. Hence,

. , (5.36)

where ∑ ; wi = gi + nqi ; gi is the dead load, qi is the live load and n is the live

load reduction factor for dynamic mass.

Live load is reduced by the n factor in calculating the lateral earthquake forces which

are based on the weight (mass) of the building during the earthquake. It is considered that the

entire live load assumed in gravity design (factored dead load and live load combination) has

a small probability of existence during an earthquake. Accordingly the mass of the live load is

reduced by n < 1 to prevent overdesign. This factor mostly depends on how long the live load

mass can be permanent in a building. It is 0.3 for residences and offices, 0.6 for schools,

dormitories, concert halls, restaurants and shops.

Page 112: Ce 490 Lecture Notes-combined

112

Lateral Force Distribution in the Code The distribution of lateral forces to the stories, given in Eq. (5.35), is slightly modified

in the Code and expressed by Eq. (5.37).

∆ ∑

; ∆ 0.0075 (5.37)

ΔFN is the portion of lateral force applied to the roof in order to produce higher mode

response. This is not an effective way of including higher mode response however. It only

leads to larger overturning moments at the base.

The lateral force distribution in the Code for the Equivalent Static Lateral Load

Procedure is shown in Fig. 5.14.

Fig. 5.14. Earthquake force distribution in the Code for equivalent static lateral load procedure.

Example 5.4. Determine the equivalent static lateral load distribution for the 3-story frame in Example 3.6. Let R=1 for a consistent comparison with the modal force vectors obtained in Example 3.7.

Solution The total weight W is 5mg where m=175 tons. Sae(T1)=1.0g since T1=0.40 s. Also, Ra(T1)=1.0. When these terms are substituted into Eq. (5.36), base shear is calculated as;

Vb=8,583.75 kN and FN = 0.0075NVb =193.13 kN. Let Hi = ih for i =1-3. Also wi =mi g. Then, from Eq. (5.37),

8,390.6229 1,864.58

8,390.6249 3,729.16

8,390.6239 ∆ 2,796.87 193.13 2,990

When these results are compared with the results of Example 3.7, it can be observed that the results of equivalent static lateral load procedure are 17.7% higher than that of mode superposition procedure. This is considered as a trade off for using a simpler procedure.

Page 113: Ce 490 Lecture Notes-combined

113

Estimation of the First Mode Period T1

The first mode period T1 is required in calculating the reduced base shear force from

Eq. (5.36). T1 can be calculated by an approximate procedure, called Rayleigh’s method. Let

Ff be a lateral force distributions and df be the resulting lateral story displacements, obtained

by static analysis. Then

2 ∑ ∑

/ (5.38)

Ff can have any distribution, however using in Eq. (5.35) for Ff increases accuracy.

Limitations of the Equivalent Static Lateral Load Procedure

Equivalent static lateral force procedure is only applicable to regular buildings where

the first mode in the earthquake direction is dominant on earthquake response. The limitations

are;

HN ≤ 25 m and ηb ≤ 2.0

in seismic zones 1 and 2, where HN is the total building height from the base, and

η ∆ ,∆ ,

(5.39)

is the torsional irregularity coefficient defined in Fig. 5.15 for the i’th story floor slab.

Fig. 5.15. Definition of the terms in torsional irregularity coefficient.

Page 114: Ce 490 Lecture Notes-combined

114

5.4.4 Accidental Torsion

The centre of rigidity of the original structure at any floor may be shifted because of

several reasons, and accordingly creates an additional eccentricity which is not accounted for

in the structural modeling of the system. This phenomenon is called accidental torsion in

seismic design. The probable causes of accidental torsion are;

• Non-synchronized cracking or yielding of lateral load resisting vertical members

(columns and shear walls) in a story, leading to unsymmetrical stiffness loss during a

strong earthquake excitation.

• Unsymmetrical distribution of non-structural members which carry a part of the story

shear forces (partition walls, window frames, etc.).

• Shift in the centre of rigidity due to imperfections in construction.

• Shift in the centre of mass due to a concentrated live load mass.

Seismic design codes account for accidental torsion indirectly, by imposing an additional

eccentricity to the applied inertial lateral force resultants. This is achieved by shifting the

centre of mass at each story in the direction perpendicular to the earthquake direction by an

additional eccentricity equal to 5% of the floor dimension in the transverse direction. The shift

is imposed in the x and y directions separately, and in both (+/-) senses. The shift at the i’th

floor in the x direction is shown in Fig. 5.16.

Fig. 5.16. Shifting the centre of mass to account for accidental torsion.

REFERENCES Turkish Earthquake Design Code for Buildings. Ministry of Reconstruction, Ankara, 2007.

Page 115: Ce 490 Lecture Notes-combined

115

6.

Seismic Design Principles for RC Structures

6.1 INTRODUCTION

The basis for the design of RC members and systems in Turkey is the standard TS-500

which is the compulsory concrete design standard under non-seismic conditions. Seismic

design rules for reinforced concrete structures are implemented by imposing additional

requirements to the conventional design under non-seismic loads. Additional requirements for

the seismic design of RC members and systems are given by Chapter 3 of the Turkish

Earthquake Code.

Seismic design of structural systems is essentially based on an inherent ductile

response under strong ground excitation. Ductility can be defined as the capacity of

undergoing large plastic deformations without reduction in strength, both at the material,

component and system levels. Although ductility is not expressed explicitly in the seismic

analysis procedures recommended by seismic design codes, reduction of elastic earthquake

forces discussed in Section 5.3 relies on the premise of a ductile seismic response.

Ductility in RC systems is considered in the global level in the Turkish Earthquake

Code. It is described for three system categories with the related seismic response

characteristics, expressed with the associated R factors.

a) Systems with enhanced ductility (R = 6‐8 in Table 2.5, TEC 2008)

b) Systems with ordinary ductility (R = 4 in Table 2.5, TEC 2008)

c) Systems with mixed ductility (R ≈ 5‐7 in Table 2.5, TEC 2008)

Systems with ordinary and enhanced ductility are essentially frame or frame-wall

systems where both the frames and the walls have either ordinary or enhanced ductility,

respectively. Systems with mixed ductility are frame-wall systems composed of frames with

ordinary ductility and walls with enhanced ductility. This is practical and leads to more

Page 116: Ce 490 Lecture Notes-combined

116

economical designs in seismic zones 3 and 4. Enhanced or mixed ductility definitions should

be satisfied in both earthquake directions.

Systems with both enhanced and ordinary ductility levels are composed of members

with similar longitudinal and lateral reinforcement arrangement, and same confinement and

anchorage requirements at critical sections in order to ensure a minimum required ductile

response under design earthquake forces. This is considered as equivalent to R = 4.

Special additional detailing rules are applied for enhanced ductility, leading to larger

reduction in elastic forces (R = 6 ‐ 8). Enhanced ductility level in RC systems which are

composed of ductile beams, columns and shear walls and strong connections is provided by

employing the “Capacity Design Principles”.

Capacity Design Capacity design has two major implications, one at the member level, and the other at

the system level.

Member level: Flexural failure mode is ensured by suppressing shear failure (capacity

shear principle in beams, columns, walls and connections).

System level: The spreading of plastic regions that undergo flexural yielding follows a

hierarchy for obtaining a more ductile system response (strong column-weak beam

principle at the connections).

6.2 DUCTILITY IN REINFORCED CONCRETE Ductility or deformation capacity of reinforced concrete systems is provided by the

ductility of its constituent materials (steel and concrete), ductility of its members (beams,

columns and walls), and the overall ductility of the structural system under seismic actions.

These are discussed separately below. However it should be noted that a ductile reinforced

concrete response can be obtained only if the dominant failure mode of the structural

components is flexure. Therefore brittle failure modes such as shear, diagonal tension and

compression should be prevented whereas ductility in flexure should be enhanced for

obtaining a ductile system response under strong seismic excitations.

6.2.1 Ductility in Reinforced Concrete Materials A ductile flexural member behavior can be achieved by employing materials with

ductile stress-strain behavior at the critical sections where bending moments are maximum.

Page 117: Ce 490 Lecture Notes-combined

117

One of the two constituents of reinforced concrete is steel, which is inherently ductile. Typical

stress-strain relationships for structural steel are shown in Fig. 6.1 for the (a) hot rolled and

(b) cold worked reinforcing bars where the horizontal axis is strain. It is observed that both

types of structural steel have large plastic strain capacity, exceeding 12 percent.

Fig. 6.1. Stress-strain relationships for structural steel. Modulus of elasticity Es 2X105 MPa.

The other constituent material, plain concrete does not possess such ductile uniaxial

material stress-strain behavior (see curve with σ2=0 in Fig. 6.2). However when the conditions

of stress change from uniaxial (σ2 = 0) to triaxial (σ2 > 0), both stress and strain capacities of

concrete enhance significantly with the increasing lateral pressure, as shown in Fig. 6.2.

Triaxial stress state in reinforced concrete members can be provided with confinement

reinforcement. When concrete is subjected to axial stress σ1, passive lateral pressure σ2

developed by the lateral tie reinforcement (Fig. 6.3.a) provides enormous increase in the

strength and strain capacity of concrete. The improvement is strongly related with the tie

spacing “s” (Fig. 6.3.b and c).

Fig. 6.2. Stress-strain relationships for concrete under uniaxial (σ2=0) and triaxial (σ2 > 0) stress.

Page 118: Ce 490 Lecture Notes-combined

118

Fig. 6.3. Confinement of concrete in rectangular sections. a) Lateral confinement pressure provided

with lateral ties, b) confined concrete along elevation, c) confined concrete along cross section.

Strength and deformation capacities of concrete fibers in the core region of columns

increase with the amount of lateral confinement reinforcement (Fig. 6.4). Confinement is most

effective in circular columns since lateral pressure develops uniformly in all radial directions

whereas a rectangular tie is more effective at the corners as shown in Fig. 6.3.a.

Fig. 6.4. Stress-strain relations for concrete in unconfined (curve 5) and confined (curves 1-4) reinforced concrete sections.

6.2.2 Ductility in Reinforced Concrete Members

Reinforced concrete members, namely beams, columns and shear walls may either fail

in flexure by reaching the yielding at a cross section, or they may fail in shear or diagonal

tension under the applied internal forces (bending moment, shear force, axial force) during an

earthquake. Shear failure is brittle since it is accompanied with no deformation capacity once

the shear capacity is exceeded. Flexure failure on the other hand is generally ductile. After the

yield moment is attained under bending, the tension steel may continue to elongate in the

Page 119: Ce 490 Lecture Notes-combined

119

plastic range until it ruptures or the concrete in compression crushes. This leads to large

plastic rotation capacities, hence large ductility capacity at member ends where yielding

occurs.

Seismic design rules for obtaining a ductile flexural behavior in reinforced concrete

beams, columns and shear walls are discussed separately below.

6.3 SEISMIC DESIGN OF DUCTILE REINFORCED CONCRETE BEAMS

Ductility in reinforced concrete beams is sensitive to the level of shear stress, ratio of

tensile reinforcement, ratio of compression reinforcement, and ratio of lateral reinforcement

for the confinement of concrete.

6.3.1 Minimum Section Dimensions Minimum dimensions are imposed for providing the space required for reinforcing

detailing to ensure ductile flexural behavior.

• Minimum beam width is 250 mm. • Beam width is limited with the adjoining column dimension perpendicular to beam. • Minimum beam height is the largest of 3 times of the slab thickness or 300 mm. • Maximum beam height is the lesser of 3.5 times the beam width or 1/4 of the beam

clear span.

The last item controls lateral buckling phenomenon and prevents deep beam behavior which

may induce flexure-shear failure mode at large deformations.

6.3.2 Limitations on Tension Reinforcement Tension reinforcement at both support and span sections should satisfy the following

limits: . 0.02 or in TS 500 (6.1)

The minimum tensile reinforcement ratio controls cracking of concrete in service conditions

whereas the maximum ratio controls ductility of the section. Let’s consider the moment-

curvature relationships of two beam sections in Fig. 6.5 with identical lateral reinforcement

and compression reinforcement ratio that ensure flexural failure mode. The only difference is

the ratio of tension reinforcement. It can be observed that ductility reduces with increase in

tension reinforcement, and 0.02 appears as a reasonable upper limit for a ductile response.

Page 120: Ce 490 Lecture Notes-combined

120

Fig. 6.5. The effect of tension reinforcement ratio on the moment-curvature ductility of beams.

6.3.3 Minimum Compression Reinforcement The ratio of bottom reinforcement to top reinforcement at the support regions of

beams should be at least 0.5 in seismic zones 1 and 2, and 0.3 in seismic zones 3 and 4. This

is indeed the ratio of compression steel to tension steel (ρ’/ρ) at the support regions.

Compression reinforcement is known to be increasing the ductility of a beam cross

section significantly. Moment-curvature relations for beam sections with minimum tension

and lateral reinforcement, but with different compression reinforcement ratios are shown in

Fig. 6.6. It is evident that ρ’/ρ 0.5 is a reasonable limit to ensure ductility. Furthermore,

when the direction of earthquake excitation reverses which may lead to a moment reversal at

the beam support, the compression reinforcement serves as the tension reinforcement.

Fig. 6.6. The effect of compression reinforcement on the moment-curvature ductility of beams.

Page 121: Ce 490 Lecture Notes-combined

121

6.3.4 Minimum Lateral Reinforcement for Confinement The under-reinforced beams are known to be ductile under monotonically increasing

moments. However when a beam undergoes moment reversals during a strong earthquake and

plastic regions (hinges) form at the end regions, confinement of these regions are necessary in

order to prevent crushing of concrete and buckling of longitudinal bars. The absence of

confined end regions results in strength degradation (reduction of yield moment under

moment reversals), which leads to extensive damage accumulation in the yielding regions.

The length of the confined regions at both ends should not be less than twice the depth

of the beam 2hk. Special seismic ties shall be used in these regions, with spacing sk being the

lesser of 150 mm, 1/4 of the beam height, or 8 times the smallest longitudinal bar diameter.

The first tie spacing from the column face should be less than 50 mm (so <50 mm) for

capturing the first shear crack on the beam end. Detailing of the confined regions of a typical

ductile beam is given in Fig. 6.8.

6.3.5 Shear Design of Beams

Shear failure is suppressed by calculating the design shear force from flexural

capacity, but not from analysis. The shear forces due to gravity loading are also taken into

account. This is called “capacity shear” in earthquake resistant design.

Let’s consider the free body diagram of a beam carrying uniform gravity loading

(g q) along its span. When the earthquake moments act on the beam, we will assume that

both end sections i and j reach their flexural capacities (with strain hardening) under double

flexure (Fig. 6.7.a).

Fig. 6.7. a) Free body diagram of a beam under gravity loads and earthquake moments at flexural capacity, b) superposition of gravity and earthquake shear forces at beam ends.

Page 122: Ce 490 Lecture Notes-combined

122

Then the shear forces (capacity shear) acting at both ends can be calculated from Fig.

6.7.b by using the principle of superposition. From equilibrium;

, , where (6.2)

Ve is the design shear force in Eq.(6.2). The top and bottom plastic moment capacities Mpi and

Mpj should consider strain hardening of tension steel. This corresponds to a 40% increase in

the yield moment capacities on average. Ve is theoretically the maximum shear force that can

develop in a beam under extreme earthquake excitation which may exceed the design

earthquake intensity (which was reduced with R). The shear capacity Vr supplied in design

should exceed the capacity shear force Ve which requires

(6.3)

The shear capacity of the column section Vr is calculated according to TS 500.

Further, possible compression strut failure due to over-reinforcing in shear is prevented by the

requirement

0.22 (6.4)

Reinforcement detailing of a typical beam designed according to TEC 2008

requirements is shown in Fig. 6.8.

Fig. 6.8. Detailing of a ductile beam designed according to TEC 2008 (dimensions in cm).

Page 123: Ce 490 Lecture Notes-combined

123

6.4 SEISMIC DESIGN OF DUCTILE REINFORCED CONCRETE COLUMNS

Ductility in reinforced concrete columns is sensitive to the level of axial load, ratio of

longitudinal reinforcement, ratio of lateral reinforcement for the confinement of concrete and

the shear forces acting on the column. Each parameter is governed with a different design

rule.

6.4.1 Limitation on Axial Stresses Moment-curvature relationship for reinforced concrete column sections is highly

sensitive to the level of axial load. The moment capacity and stiffness increases with axial

load up to N/No 0.4, but ductility decreases as shown in Fig. 6.9. Therefore the level of

axial load or axial stress has to be limited such that ductile section response is ensured.

Fig. 6.9. The effect of axial load on the moment-curvature response of rectangular columns.

There are two requirements for limiting axial stresses in columns. The first one

controls the minimum cross section dimensions in rectangular columns:

75 000 ; 250 (6.5)

The second one controls the maximum axial stress in columns.

0.50 ; 20 (6.6)

6.4.2 Limitation on Longitudinal Reinforcement A ductile flexural column response requires that the minimum and maximum

longitudinal reinforcement ratios are 1% and 4%, respectively.

0.01 0.04 (6.7)

Page 124: Ce 490 Lecture Notes-combined

124

6.4.3 Minimum Lateral Reinforcement for Confinement Lateral ties which confine concrete in compression increases the compressive strain

capacity of concrete as discussed in Section 6.2. This increase in the strain capacity of

concrete also improves the curvature ductility of the cross section significantly. Moment-

curvature relationships of two typical rectangular column sections are shown in Fig. 6.10.

Two columns are identical except the amount of lateral reinforcement. The lower curve is for

a column with inadequate lateral reinforcement, which represents typical non-seismic design

detailing. It has an extremely brittle response with almost no deformation capacity after

yielding. The upper curve is for a column with minimum lateral reinforcement according to

TEC 2008. Even minimum lateral reinforcement improves the curvature ductility of the

column cross section enormously.

Fig. 6.10. The effect of confinement on the moment-curvature response of rectangular columns.

Minimum lateral reinforcement is required for the end regions of columns where

yielding is expected to occur. These regions are called the confinement regions at both ends

and their length should not be less than the larger of 500 mm, 1/6 of the column clear length,

or the larger dimension of the column section.

Minimum diameter of the tie reinforcement is 8 mm and the minimum vertical spacing

s is the lesser of 100 mm or 1/3 of the smaller section dimension. The maximum lateral

distance a (see Fig. 6.3.a and 6.17) between the legs of tie reinforcement should be less than

25 tie diameters. Cross ties are added to satisfy this requirement if necessary.

The minimum area of lateral reinforcement should satisfy the larger value in Eq. (6.8)

for Nd 0.20 Ac fck. If Nd is less, 2/3 of the value calculated from Eq. (6.8) is used.

0.30 1 or 0.075 (6.8)

Page 125: Ce 490 Lecture Notes-combined

125

The parameters in Eq.(6.8) are explained on Fig. 6.17. The minimum lateral reinforcement in

Eq. (6.8) maintains the flexural strength of a column after cover spalling, following the

formation of a plastic hinge at the considered section.

6.4.4 Strong Column-Weak Beam Principle

The system level implementation of capacity design was manifested in Section 6.1 as

“the spreading of plastic regions that undergo flexural yielding follows a hierarchy for

obtaining a more ductile system response”. Flexural plastic hinges inevitably form at the ends

of frame members under design ground motions which are reduced with R > 1.

A plastic hinge formed on a beam is less critical than a plastic hinge on a column or

shearwall, because vertical members may loose their stability under gravity loads when plastic

hinges form. Accordingly, plastic hinge hierarchy requires formation of plastic hinges first on

beams, then at the base of first story columns. A plastic hinging hierarchy can be imposed in

design by proportioning the flexural capacities of beam and column ends joining at a

connection. This is called the strong column-weak beam principle, which is expressed by

Eq.(6.9) and explained in Fig. 6.11. The constant 1.2 is a safety factor in Eq. (6.9).

, , 1.2 , , (6.9)

Fig. 6.11. Moment capacities of member ends joining at a beam-column connection.

Strong column-weak beam principle leads to a ductile collapse strategy under

increasing lateral earthquake forces. Let’s consider the three story, single bay frame in Fig.

6.12.a, which obeys strong column-weak beam design. It is loaded with increasing lateral

forces until collapse. This type of inelastic analysis is called the pushover analysis in

earthquake engineering. If we plot the base shear force versus the roof displacement, we

obtain the capacity curve in Fig. 6.12.b. The progress of plastic hinges at different stages of

loading are marked on both figures. It can be observed that the frame exhibits significant

ductility before collapse under lateral forces.

Page 126: Ce 490 Lecture Notes-combined

126

Formation of plastic hinging at the column (or wall) bases is inevitable at the later

stages of loading since they become cantilever columns after all the beams spanning to the

column yield, as shown in Fig. 6.12.c.

Fig. 6.12. (a) Plastic hinge hierarchy in a frame with strong columns and weak beams under increasing lateral loads. (b) Capacity curve of the frame in (a). (c) Free body diagram of a column when all connecting beams yield. Further increase in lateral loads will lead to yielding at the column base.

If columns are weaker than beams, then a soft story may develop. Plastic hinges form

at the first story columns first where the moments are maximum (Fig. 6.13.a). When plastic

hinges develop both at the bottom and top ends of the first story columns, a soft story forms

since the instantaneous lateral stiffness becomes very low, even zero when there is no strain

hardening (Fig. 6.13.b). Then lateral deformations increase quickly under increasing lateral

loads, and the frame looses its stability under gravity loads. This is a catastrophic collapse

(pan caking). Most collapses and life losses in buildings during strong earthquakes are due to

soft story formation.

Fig. 6.13. (a) Plastic hinge hierarchy in a frame with weak columns and strong beams under increasing lateral loads. (b) Capacity curve of the frame in (a).

Page 127: Ce 490 Lecture Notes-combined

127

Two examples of soft story damage and collapse are shown below in Fig. 6.14.

Fig. 6.14. Soft story collapses from (a) 1971 San Fernando Earthquake, USA on the left, (b) 2003 Bingöl earthquake, Turkey on the right.

6.4.5 Shear Design of Columns

The design shear force is also calculated from capacity shear, but not from analysis as

discussed in the shear design of beams. There is no span loading however on columns. Let’s

consider a column which reaches its flexural capacity (with strain hardening) at both top and

bottom ends under double flexure (Fig. 6.14.a). From equilibrium;

, , (6.10)

where Ve is the design shear force, Mp,top and Mp,bot are the plastic moment capacities with

strain hardening at the top and bottom ends, respectively.

Fig. 6.14. A column under double flexure. (a) Yielding at column ends: weak column-strong beam, (b) yielding at beam ends: strong column-weak beam.

Page 128: Ce 490 Lecture Notes-combined

128

Eq. (6.10) is valid when columns are weaker than beams. In capacity design however,

columns are stronger than beams. Hence beam ends yield before column ends around a

connection (Fig. 6.14.b). In this case, equilibrium of moments around the connection requires

~ , , (6.11)

Hence, the sum of plastic column end moments in Eq.(6.10) is replaced with the right hand

side of Eq.(6.11). Clearly, this replacement reduces the column design shear force in

Eq.(6.10).

The shear capacity Vr supplied in design should exceed the capacity shear force Ve.

(6.12)

The shear capacity of the column section Vr is calculated according to TS 500. Possible

compression strut failure due to over-reinforcing in shear is prevented by the requirement

0.22 (6.13)

where Aw is the shear area in the earthquake direction.

6.4.6 Short Column Effect Short columns may lead to brittle shear failure even if the original column is designed

for capacity shear. Short columns form due to shortening of clear length from to ′ by

architectural interventions such as parapet walls (Fig. 6.15). Then the shear force which

develops in the short column ′ can be estimated from Eq.(6.14) by replacing with ′ in

Eq.(6.10).

,

(6.14)

is the plastic moment capacity of the column at the bottom of clear length. Since ′ ,

naturally ′ . Such increase in the shear force usually reverses the design inequality in

Eq. (6.12) which causes shear failure during an earthquake.

Furthermore, the lateral stiffness of a short column is much larger than the original

column.

′ ′ (6.15)

Page 129: Ce 490 Lecture Notes-combined

129

Accordingly, a short column attracts larger internal shear forces compared to its original state

even if Mp is small in Eq.(6.14), and reaches its shear capacity easily.

Short column formation can be prevented by separating the parapet walls or

architectural obstacles from the columns which reduce their effective clear lengths.

Otherwise, short columns which develop intentionally should be designed for the revised

shear force calculated from Eq.(6.14).

Fig. 6.15. Short column formation.

A short column failure during the 1999 Düzce earthquake is shown in Figure 6.16.

Fig. 6.16. Short column failures in a school building after the 1999 Düzce earthquake. Picture on the right shows detailed shear damage from one of the short columns.

Reinforcement detailing of a typical column designed according to TEC 2008

requirements is shown in Fig. 6.17. Field application of column reinforcement is also shown

at the inset of the same figure.

Page 130: Ce 490 Lecture Notes-combined

130

Fig. 6.17. Detailing of a ductile column designed according to TEC 2007, and the photograph of a

column reinforcement cage in the field (dimensions in cm).

 

Page 131: Ce 490 Lecture Notes-combined

131

6.5 SEISMIC DESIGN OF BEAM-COLUMN JOINTS IN DUCTILE FRAMES

A beam-column joint is the part of the column where the beams join. Joints are

considered as brittle components in seismic design. Accordingly, shear strength of a joint

should be larger than the maximum shear force acting on the joint during an earthquake for an

acceptable design. A typical interior joint is shown in Fig. 6.18.

Beam-column joints in a ductile frame are classified into two types in seismic design.

If beams are connecting to the joint from all four sides, and if the width of these beams is at

least ¾ of the width of the column that they are joining to, then such a joint is classified as a

confined joint (bw1 and bw2 both larger than 3/4 b; bw3 and bw4 both larger than 3/4 h in Fig.

6.18.b). Only some interior joints can satisfy this condition. All other joints which do not

satisfy the above condition are classified as unconfined joints. The contribution of concrete to

joint shear strength is larger in confined joints compared to unconfined joints, as given in

paragraph 6.5.2 below.

Fig. 6.18. A typical beam-column joint. (a) Equilibrium of horizontal (shear) forces acting on an

interior joint, (b) confinement conditions of a joint determined by the spanning beams.

As1

As2

Vbot

Vcol = min (Vtop , Vbot)

Vtop

1.25As2fyk 1.25As1fyk C1

C2

(b)

bw1

b

bw2

bw3

bw4 h

b2 b1

Earthquake direction

(a)

Page 132: Ce 490 Lecture Notes-combined

132

6.5.1 Design Shear Force Design shear force acting along a joint can be calculated with the aid of Fig. 6.18.a.

The design shear is not calculated from analysis, but it is calculated from the flexural

capacities of the beams spanning into the joint according to capacity design principles. Since

the frame design is based on the strong column- weak beam principle, it is assumed that

maximum shear force in the joint develops with the flexural yielding of beams in the

earthquake direction connecting from both sides. These beam ends are in opposite bending

directions under lateral earthquake effects, hence top longitudinal reinforcement on the left

beam end and bottom longitudinal reinforcement on the right beam end yields under the

bending directions shown in Fig. 6.18.a. If a horizontal cross section is taken from the middle

of joint in Fig. 6.18.a, we can calculate the joint shear along this section from equilibrium.

Considering that the compression in concrete is equal to tension in reinforcement at a section

in bending, the design shear force simply becomes

1.25 + )- (6.16)

where is the characteristic yield strength of longitudinal reinforcement and is the

smaller shear force in the connecting columns at top and bottom. The smaller value is taken

because column shear acts against the shear imposed by the beams. 0 when the joint is

at the exterior frame.

6.5.2 Design Shear Strength The joint design shear strength is given in Equations (6.17.a) and (6.17.b) for

confined and unconfined joints, respectively.

0.60 (6.17.a)

0.40 (6.17.b)

It can be noticed that the contribution of joint lateral reinforcement to joint shear

strength is not given explicitly in Eqs. (6.17). Lateral reinforcement requirements in joints are

controlled by the minimum reinforcement conditions. In confined joints, 40 % of the column

lateral reinforcement at the column confinement region should be placed in the joint region,

however maximum spacing is 150 mm. These figures are 60 % and 100 mm for unconfined

joints. More amount of horizontal reinforcement is placed in unconfined joints since the

Page 133: Ce 490 Lecture Notes-combined

133

contribution of concrete is less, as discusses above. The minimum horizontal bar diameter is 8

mm in both types of joints. Details of a confined and an unconfined joint reinforcement are

shown in Fig. 6.17.

A joint design is acceptable for shear if

(6.18)

Otherwise the width of the column (hence the size of the joint) or the depth of the beam (for

reducing longitudinal reinforcement) should be increased.

An example of joint failure is shown in Fig. 6.19. Joint shear strengths were too low to

resist the shear forces transmitted by beams.

Fig. 6.19. Beam-column joints failed during the 2011 Van earthquake. 6.6 SEISMIC DESIGN OF DUCTILE CONCRETE SHEAR WALLS

Shear walls mainly serve as the major component of a building for resisting the lateral

loads applied during an earthquake. Their lateral stiffnesses compared to typical columns are

very large due to their large size in the direction of earthquake forces. A typical shear wall is

shown in Figure 6.20. A large column qualifies as a shear wall if length to thickness ratio of

its cross section is larger than 7. The minimum thickness is either 200 mm or 1/20 of the story

height. In buildings where the entire lateral load resisting system are composed of shear walls

(tunnel form buildings in particular), minimum thickness may be reduced to 150 mm. The

thickness of a shear wall may be reduced along the wall height, depending on the design

requirements.

Page 134: Ce 490 Lecture Notes-combined

134

7 Frame-wall buildings 200

Tunnel form buildings 150

Fig. 6.20. A typical shear wall and minimum dimensions.

Shear walls are classified into two types regarding their deformation behavior. The

classification is based on the height to length ratio, or the aspect ratio of the wall.

2.0 : Slender walls - Flexural behavior governs.

2.0 : Squat walls - Shear behavior governs.

The deformation behavior of slender (flexure dominant) and squat (shear dominant)

walls are shown in Fig. 6.21. Slender walls deform similar to cantilever columns under lateral

forces, hence flexural deformations dominate wall deformations. A slender wall reaches its

lateral strength capacity with flexural yielding at its base region, with the formation of

horizontal flexural cracks accompanied with the yielding of longitudinal tension steel (Fig.

6.21.a). Shear deformations dominate deformation of squat walls. A squat wall reaches its

lateral strength capacity with the formation of a family of diagonal cracks under diagonal

tensile stresses, accompanied with the yielding of web reinforcing bars which are crossing

these cracks (Fig. 6.21.b).

Fig. 6.21. Deformation behaviors of (a) a slender shear wall, (b) a squat shear wall.

Page 135: Ce 490 Lecture Notes-combined

135

6.6.1 Seismic Design of Slender Shear Walls

The behavior of slender walls is similar to ductile columns except the bending moment

distribution (Fig. 6.22). A column bends in double curvature under lateral forces and reaches

its lateral capacity with the formation of plastic hinges forming at the beam ends connecting

to the column when the columns is stronger than the beam (Figs. 6.14.b and 6.22.a). Slender

shear walls on the other hand usually bend in single curvature because the flexural capacities

of the connecting beams are too small to change the sign of moment distribution along the

wall (Fig. 6.22.b). Maximum bending at the base cannot exceed the plastic moment capacity

Mp, base of the base section.

(a) (b)

Fig. 6.22. Moment and curvature distributions along (a) a column, (b) a slender shear wall.

Reinforcement Detailing in Ductile Shear Walls

Ductility capacity of a shear wall is therefore controlled by the ductility capacity of its

plastic zone at the base. The height of this zone is Hcr where,

, 2

Along Hcr, confined end regions are provided at the edges of the wall section at each end

region (Fig. 6.22). The minimum length of the confined end regions in plan along Hcr are

defined as; 2 0.20 (larger governs). Confined end regions are also

formed above Hcr, but their length is reduced to half, i.e. 0.10 (larger

governs). The ratio of longitudinal reinforcement to the total wall area ( As/Aw) within

“each” confined end region is 0.002 along Hcr and 0.001 above Hcr. The details of lateral

confinement reinforcement at the confined end regions are similar to the details given for

column confinement region in Section 6.4.3.

Page 136: Ce 490 Lecture Notes-combined

136

The reason for confining the end regions of walls along the plastic hinge regions is

similar to that of columns, but also this is more economical for the effectiveness of vertical

flexural reinforcement. When ultimate moment develops at a wall section, the vertical

reinforcement in the confined region at the tension side is much more effective than the

vertical reinforcement distributed along the web because of its lever arm from the neutral axis

(Fig. 6.23). On the other hand, the end region at the compression side is subjected to high

compressive stresses and strains. Confining concrete in this region increases the compressive

strain capacity of concrete, accordingly it increases the curvature capacity of the section long

the critical height Hcr where wall plastic hinge develops.

Uniform layers of horizontal and vertical reinforcement are placed in the web between

the confined end regions, at both faces. The ratio of minimum web reinforcement is 0.0025 in

both vertical and horizontal reinforcement, with a maximum spacing of 250 mm.

Fig. 6.23. Strain distributions along the wall length within Hcr and reinforcement detailing.

Page 137: Ce 490 Lecture Notes-combined

137

Flexural Design of Slender Shear Walls

Design moment distribution in ductile slender walls under earthquake loading is

defined in Fig. 6.24. Moment diagram obtained from analysis usually reaches its maximum at

the base and decreases with height. In wall-frame systems, moment diagram changes sign at

upper stories because of the interaction of the wall with the frame. When beams are spanning

into the shear wall, moment diagram displays discontinuities at the floor levels due to the

resisting moments of the beams with opposite sign, however these moments are very small

compared to the wall moments and they can be ignored in the wall moment distribution.

Design moments are obtained from analysis moments by drawing a chord from the

base moment to the zero moment at the wall top, and then shifting this chord up by the

distance Hcr as shown in Fig. 6.24. The reason for increasing the design moments at the

support region (along Hcr) is the increase of tensile forces due to tension shift.

Fig. 6.24. Design moment distributions along the wall length

Tension shift is essential in the curtailment of longitudinal reinforcement in deep

beams and shear walls. Let’s consider the cantilever shear wall in Fig. 6.25.a under the effect

of lateral forces. Horizontal reinforcement (stirrups) at the web is activated with the formation

of a diagonal shear crack, and ultimate capacity of the shear wall is reached with the yielding

of tensile steel. A free body diagram of the wall above the shear crack is shown in Fig. 6.25.b.

where T2 is the tension in longitudinal steel at section 2, C1 is the compression at section 1, z

is the moment arm, Vs is the shear at stirrups crossing the diagonal crack and Vb is the base

shear at section 1. Moment equilibrium about the base requires

Analysis Bending moment 

Design bending moment  

Shear wall – frame sytem Shear wall system

HwHw

HcrHcr

Analysis Bending moment 

Design bending moment  

Page 138: Ce 490 Lecture Notes-combined

138

0.5 (6.18)

where M1 is the internal resisting moment at section 1, or at the base. Also, from the

equilibrium of the vertical cantilever part 1-2 in Fig. 6.25.d,

(6.19)

Here, V2 and M2 are the internal shear and moment at section 2, respectively and F1‐2 is the

portion of external lateral force acting on the wall between 1-2. T2 can be obtained from Eqs.

(6.18) and (6.19).

0.5 0.5 0.5 0.5 (6.20)

In Eq. (6.20), 0.5Vs is small compared to (V2 + 0.5F1‐2) where Vs is the force in the stirrup

crossing the diagonal crack whereas V2 is the total shear at section 2. Then it can be ignored.

Accordingly, Eq. (6.20) reduces to

(6.21)

with the aid of Eq. (6.19). Hence, tension at section 2 is not related to the moment at section 2

as in a homogeneous, uncracked wall but it is calculated from the larger moment at section 1.

Fig. 6.25. (a) External forces acting on the wall, (b) internal forces acting along the diagonal crack, (c) moment diagram of the wall.

Page 139: Ce 490 Lecture Notes-combined

139

The moment capacity of a shear wall can be simply calculated by ignoring the

contribution of vertical reinforcement in the web by using Eq. (6.22), with the aid of Fig.

6.23.

1

1 (6.22)

where As is the area of vertical reinforcement in the confined region at the tension side, c is

the depth of neutral axis, Nd  is the axial compressive force acting on the wall section under

vertical loads and fy is the yield strength of vertical steel. Mr should exceed the design

moments calculated according to Fig. 6.24 at all sections. The moment capacity of shear walls

should also satisfy the strong column-weak beam principle defined for columns in paragraph

6.4.4.

Shear Design of Slender Shear Walls

Design shear in ductile walls is calculated from analysis, but it is modified by the

plastic moment capacity of the base.

. (6.23)

Here, Md and Vd are the moment and shear calculated from analysis under vertical loads and

linear elastic earthquake forces (Ra=1), and is the dynamic magnification factor for taking

into account the higher mode effects. Its value is 1.5 for frame-wall systems. At the wall base,

Md cannot exceed the plastic moment capacity Mp. Hence the shear force distribution

calculated from analysis is scaled by at the wall base. If detailed analysis is not

conducted, then ~ 1.25 can be assumed. Further, Ve cannot exceed Vd calculated from

analysis under vertical loads and earthquake forces reduced with Ra=2.

2 (6.24)

An acceptable shear design should satisfy for diagonal tension, and 0.22

for diagonal compression where

0.65 (6.25)

Here, and are the compressive and tensile design strength of concrete, is the ratio

of horizontal web reinforcement and is the design strength of horizontal reinforcement.

Page 140: Ce 490 Lecture Notes-combined

140

An example of a slender wall detailing which is designed according to the 2007 TEC

is shown in Fig. 6.26 below.

Fig. 6.25. Detailing of a ductile slender shear wall designed according to TEC 2007.

 

Page 141: Ce 490 Lecture Notes-combined

141

6.6.2 Seismic Design of Squat Shear Walls

Squat walls cannot exhibit a ductile response. Hence, Hcr and confined end regions are

not defined for these walls. Design moments are calculated from analysis. However bending

is never critical for squat walls since their moment capacity is always much larger than the

design moments.

Design shear is calculated form analysis with Ra=1, but not modified with capacity as

in the slender walls. Hence 1 . Then an acceptable shear design is achieved if

for diagonal tension, and 0.22 for diagonal compression where is given

by Eq. (6.25).

A squat shear wall which failed during the 2011 Van Earthquake is shown in Fig. 6.26.

The main reasons of this wall failure is low concrete compression capacity and insufficient

shear reinforcement, as well as out-of-plane bending moments resulting from the torsional

response of the unsymmetrical structure during the earthquake.

Fig. 6.26. A squat shear wall failed during the 2011 Van Earthquake. 6.7 CAPACITY DESIGN PROCEDURE

A summary of the capacity design procedure is given below in an algorithmic form.

Then a case study is presented for the seismic design of a 5 story concrete frame.

1. Analyze the system under , assuming flexural behavior

2. Design the beams for flexure under the analysis moments Md and design for shear

under capacity shear Ve.

Page 142: Ce 490 Lecture Notes-combined

142

3. Calculate column moments from strong column-weak beam inequality. Perform

flexural design under these moments and shear design under capacity shear.

4. Design shear walls for flexure under design moment distribution with tension shift,

and design for shear under capacity shear Ve.

5. Design the capacity of joints for capacity shear.

Example 6.1. Seismic Design of a 5 Story Reinforced Concrete Frame

Typical floor plan of a 5-story reinforced concrete frame building and all member dimensions are given below. Floor plan and member dimensions are the same in all stories. The building is located in Seismic Zone 1, and soil condition is Z3. Design Frame 3 according to the Turkish Seismic Code 2007. Concrete grade is C25 and steel grade is S420. Ductility level is “enhanced” (R=8). The earthquake direction is Y.

Plan View:

(Units in centimeters)

Page 143: Ce 490 Lecture Notes-combined

143

Frame-3 Elevation View:

(Units in meters)

Solution Story masses

Columns are 50X40 cm; beams are 30X50 cm; slab thickness is 12 cm. Finishing on slabs 1.5 kN/m2 (except the roof slab). Live load on slabs is 2 kN/m2 (according to TS498).

Dead loads (DL), live loads (LL) and story masses which are calculated consistently with the gravity loads (TEC 2007, Eqn. 2.6) are given in the table below.

Eigenvalue analysis

Eigenvalues (modal periods) and mass normalized eigenvectors (mode shapes) are calculated from eigenvalue analysis, by using the structural model in the Y direction.

Mode  1  2  3  4  5 

Period (s)  0.505 0.160 0.089 0.060  0.046 

Dead Load (kN)  Live Load (kN) Story Masses (tons) (DL+0.3LL) Columns  Beams  Slab  Finishing Slab 

60.0  125.6  202.2  101.2  135  1st Story  53,7 60.0  125.6  202.2  101.2  135  2nd Story  53,7 60.0  125.6  202.2  101.2  135  3rd Story  53,7 60.0  125.6  202.2  101.2  135  4th Story  53,7 30.0  125.6  202.2  0  135  5th Story  42,9 

Page 144: Ce 490 Lecture Notes-combined

144

0

1

2

3

4

5

‐0,1 ‐0,05 0 0,05 0,1

Stor

y #

1st Mode 2nd Mode 3rd Mode

Mass Normalized Eigenvectors Mode 1  Mode 2  Mode 3  Mode 4 Mode 5‐0,0173  0,0502  ‐0,0753  ‐0,0823 0,0579‐0,0418  0,0840  ‐0,0427  0,0446 ‐0,0774‐0,0636  0,0526  0,0668  0,0383 0,0767‐0,0794  ‐0,0217  0,0497  ‐0,0802 ‐0,0542‐0,0882  ‐0,0852  ‐0,0725  0,0495 0,0237

Response Spectrum Analysis and Equivalent lateral Forces a) Minimum number of modes

Mode  Mass Participation Ratio (%)  Cumulative  (%) 1  83,2 83,22  10,5 93,73  4,0 97,74  1,8 99,55  0,5 100,0

First two modes are sufficient for the response spectrum analysis according to TEC 2007.

b) Modal forces and base shear forces

Modal Forces (kN) Mode 1  Mode 2  16,64  17,24 40,34  28,84 61,34  18,05 76,57  ‐7,46 67,93  ‐23,36 

c) Equivalent lateral forces (TEC 2007)

We will also calculate the equivalent lateral forces for comparison with modal forces.

Total Weight= 2547,26 kN Period= 0,505 seconds S(T)= 1 A(T)= 9,81 m/s2 Base Shear = 318,40 kN for R=8 Check = 101,89 (0,1*A0*I*W) < 318.4 OK. ΔFn = 11,94 kN

Base Shear Forces (kN) 

Mode 1 Mode 2  VtB (kN) (SRSS)262,83 33,31  265.05

Page 145: Ce 490 Lecture Notes-combined

145

Equivalent lateral force procedure gives 20% larger forces compared to response

spectrum analysis.

d) Check according to TEC 2007, Section 2.8.5 VtB = 265.05 kN (Calculated from response spectrum analysis)

Vt = 318.40 kN (Calculated according to TEC2007, equation 2.4) β = 0.8. Check: . . Therefore, there is no need for base shear correction in mode superposition analysis.

Design

Response spectrum analysis is conducted under the modal forces given above and member forces and displacements are obtained by SRSS combination. Also member forces are obtained with the equivalent static lateral load method for comparison. The most critical members according to the internal forces acting on them (DL + LL EQ/R) calculated by response spectrum analysis are designed.

Story  Fi (kN) 1  21,90 2  43,80 3  65,70 4  87,59 5  99,41 

0

1

2

3

4

5

‐25 25 75 125

Stor

y #

Modal Forces and Equivalent Lateral Forces

1st Mode2nd ModeELF

Page 146: Ce 490 Lecture Notes-combined

146

Beam Design

Beam Moment and Shear Diagrams from Equivalent Lateral Load Analysis: The most critical member: K224 (Second Story, Middle Bay Beam)

(Lateral forces are applied in both +/- directions, envelope moment diagram is presented) Moment and Shear Diagrams from Response Spectrum Analysis: The most critical member: K224 (Second Story, Middle Bay Beam) Envelope Moment and Shear diagrams are presented.

Page 147: Ce 490 Lecture Notes-combined

147

Flexural Design of the most critical beam (Slab contribution is ignored) i End j End

Positive Direction Md (kN.m)= 106,175 106,173

Asreq (mm2)= 685 685 Asmin= bwd(0.8*fctd/fyd)

(TS 500 Eqn. (7.3)) Asmin (mm2)= 348 348

Negative Direction Md (kN.m)= 74,16 74,16

Asreq (mm2)= 455 455 Asmin= bwd(0.8*fctd/fyd)

(TS 500 Eqn. (7.3)) Asmin (mm2)= 348 348

Reinforcements provided Support Span

Astop (mm2) 3φ14 + 2φ16 (864 mm2) 3φ14 (462 mm2)

Asbottom (mm2) 3φ16 (604 mm2) 3φ16 (604 mm2)

Mr (kN.m) 134,96

95,8

Shear Design of the most critical beam

i End j End

(TEC 2007 Section 3.4.5) Vdy (kN) 25,72 (from gravity load analysis)

(Mp 1.4Mr) (+) Mp (kN.m) 188,95 134,17

(Mp 1.4Mr) (-) Mp (kN.m) 134,17 188,95

Positive Direction Ve (kN) 133,72

Negative Direction Ve (kN) 131,15

Shear Reinforcement provided Support Span

Asw 8φ/110 mm 8φ/200 mm

Shear Capacity of the Beam

Vw = (Asw/s)*fywd*d (TS 500 Eqn. (8.5))

Vw (kN) 85,84

Vc = 0.8*(0.65*fctd*bw*d(1+γ(Nd/Ac))) (TS 500 Eqns. (8.1 & 8.4)) Vc (kN) 103,29

Vr  = Vc  + Vw (TS 500 Eqn. (8.3))

Vr (kN) 189,13

(TEC 2007 Eqn. (3.10)) 0,22 bw d fcd (kN) 522,5

Ve < Vr ; Ve < 0,22 bw d fcd - OK

106 74

74 106

134 189Mp 1.4 Mr

Page 148: Ce 490 Lecture Notes-combined

148

Column Design

The most critical member: S13 (First Story, Middle column)

Column Flexural Design (Loads on the most critical column from combinations)

Design load on the column, from analysis

Nd (kN.m) 803,56

Md (kN.m) 120,2

ρ l = 0.01 (TEC 2007 Section 3.3.2) Asreq (mm2) 2000

Asprovided (mm2) 2010,62 10φ16 Mr (kN.m) 266,2

Minimum longitudinal reinforcement ratio governs column design. Strong Column - Weak Beam Check for the most critical connection

Columns Mra (kN.m) 266,2

Mrü (kN.m) 266,2

Beams Mri (kN.m) 134,96

Mrj (kN.m) 95,8

(Mra + Mrü) ≥ 1.2 * (Mri + Mrj) (TEC 2007 Eqn. (3.3)) Check 1,92> 1,2 OK

Column Shear Design

Most critical column: Second story middle columns.

Mü calculations for the top end ΣMp (kN.m) 323,12

Mü (kN.m) 191,22

Mü calculations for the bottom end

ΣMp (kN.m) 323,12

Ma (kN.m) 158,81

Shear force on Column Ve = (Ma + Mü) / ln

(TEC 2007 Eqn. (3.5)) Ve (kN) 100,01

Shear Reinforcement provided Region End Middle

Asw 8φ/7 8φ/19

189134

189 134

Page 149: Ce 490 Lecture Notes-combined

149

Confined region length for column top and bottoms (TEC 2007 Fig. (3.3)) 500 mm

Shear Capacity of the Column

Vw (kN) 249,01 Vc (kN) 162,35 Vr (kN) 411,36

0,22 bw d fcd (kN) 733.33

Ve < Vr ; Ve < 0,22 bw d fcd - OK Beam-Column Connection Shear Check

Connection check for column S13 (First Story, Middle Column):

Confinement Check Beam Dimensions Column Dimensions

bw1 : 0.3 m b : 0.4 m bw2 : 0.3 m h : 0.5 m bw3 : 0.3 m bw4 : 0.3 m

Check (TEC 2007 Fig. (3.10)) bw1 and bw2 ≥ 3/4b (Satisfied)

bw3 and bw4 < 3/4h (Not Satisfied) Therefore the joint is unconfined.

Shear Force on the Joint (TEC 2007 Section 3.5.2.1)

Ve = 1.25 fyk (As1+As2) - Vkol (TEC 2007 Eqn.(3.11))

As1 : 864 mm2 As2 : 603 mm2 fyk : 420 Mpa Vkol: 90.5 kN

Ve = 679.7 kN

Joint Shear Force Limit (TEC 2007 Section 3.5.2.2)

Ve ≤ 0.45 bj h fcd  (TEC 2007 Eqn.(3.13) for unconfined region)

bj : 0.4 m h : 0.5 m fcd : 16.67 MPa

 Ve < 1500 kN (Satisfied)

Page 150: Ce 490 Lecture Notes-combined

150

Beam and Column Cross-Sections (units in mm)

Beam and Column Support Sections

Beam and Column Span Sections

Pushover Analysis

Frame is modeled using the design parameters and pushover analysis is conducted. Capacity curve and plastic hinge pattern is determined.

a) Capacity Curve

Page 151: Ce 490 Lecture Notes-combined

151

b) Plastic Hinge Pattern

Response History Analysis under DZC270

Frame is dynamically analyzed under DZC270 recorded during the 1999 Düzce Earthquake. Roof displacement history and plastic hinge pattern are obtained.

a) Roof Displacement History

Page 152: Ce 490 Lecture Notes-combined

152

b) Plastic Hinge Pattern