Blandon 2005

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

This article was downloaded by: [University of Saskatchewan Library]

On: 23 August 2013, At: 11:54


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Earthquake Engineering


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/ueqe20

EQUIVALENT VISCOUS DAMPING


EQUATIONS FOR DIRECT
DISPLACEMENT BASED DESIGN
a a
Carlos Andres Blandon & M. J. N. Priestley
a
European School for Advanced Studies in Reduction of
Seismic Risk (ROSE School), Via Ferrata, 27100, Pavia, Italy
Published online: 21 Dec 2010.

To cite this article: Carlos Andres Blandon & M. J. N. Priestley (2005) EQUIVALENT VISCOUS
DAMPING EQUATIONS FOR DIRECT DISPLACEMENT BASED DESIGN, Journal of Earthquake
Engineering, 9:sup2, 257-278

To link to this article: http://dx.doi.org/10.1142/S1363246905002390

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever
as to the accuracy, completeness, or suitability for any purpose of the Content. Any
opinions and views expressed in this publication are the opinions and views of the
authors, and are not the views of or endorsed by Taylor & Francis. The accuracy
of the Content should not be relied upon and should be independently verified
with primary sources of information. Taylor and Francis shall not be liable for any
losses, actions, claims, proceedings, demands, costs, expenses, damages, and other
liabilities whatsoever or howsoever caused arising directly or indirectly in connection
with, in relation to or arising out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.
Terms & Conditions of access and use can be found at http://www.tandfonline.com/
page/terms-and-conditions
Journal of Earthquake Engineering
Vol. 9, Special Issue 2 (2005) 257–278
c Imperial College Press

EQUIVALENT VISCOUS DAMPING EQUATIONS FOR


DIRECT DISPLACEMENT BASED DESIGN
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

CARLOS ANDRES BLANDON and M. J. N. PRIESTLEY


European School for Advanced Studies
in Reduction of Seismic Risk (ROSE School)
Via Ferrata, 27100, Pavia, Italy

Estimation of equivalent viscous damping is an important step in the methodology of


the direct displacement based design. Errors in the estimation of this parameter may
lead to consequent errors in the ductility demand of the designed elements. Equivalent
viscous damping estimated by Jacobsen’s approach for steady-state harmonic response
are compared with effective damping factors obtained from an iterative procedure using
time — history analyses of single-degree-of-freedom systems. In order to separate the
effects of elastic-viscous and hysteretic damping the designs and analyses were carried
without elastic viscous damping. Six hysteretic models representing a wide range of
hysteretic energy absorption characteristics and six artificial records were used in the
analysis. It is found that in general Jacobsen’s approach overestimates the values for
equivalent viscous damping. Based on the analytical results, modified design equations
for equivalent viscous damping are proposed.

Keywords: Direct displacement based design; equivalent viscous damping.

1. Introduction
Design methodologies based on forces have generally been used in the past to define
the capacity and demand of the structural systems under seismic attack. However,
it is now generally accepted that design methodologies based on displacement are
more appropriate, and can overcome inherent deficiencies of traditional force based
design. One alternative of the methodologies based on displacement is “Direct Dis-
placement Based Design (DDBD)” proposed by Priestley [1997]. In this approach,
structures are designed to achieve, rather than be limited by displacements corre-
sponding to a specified limit state.
DDBD is based on the substitute-structure concept developed by Shibata and
Sozen [1976] which represents the structure for design purposes by the secant stiff-
ness to maximum displacement response, and equivalent viscous damping repre-
senting the combined effects of elastic and hysteretic damping. Estimation of this
equivalent viscous damping factor (ξ) (EVDF) used to characterize the substitute
structure is a key parameter in this design methodology. In the past this factor has
generally been estimated based on the methodology proposed by Jacobsen [1930,
1960]. However, Jacobsen’s approach assumes complete loops of the hysteretic mod-
els, under sinusoidal excitation. It is necessary then, to analyze the accuracy of the

257
258 C. A. Blandon & M. J. N. Priestley

estimation of the viscous damping factor using this approach, for different types
of earthquakes, structural periods (effective periods), ductility levels and struc-
tural systems (hysteretic models). The fundamentals of the direct displacement
based design will not be repeated here as they are covered in a companion paper
[Pettinga and Priestley, 2005], to which the reader is referred for details.

2. Viscous Damping
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

The concept of viscous damping is generally used to represent the energy dissipated
by the structure in the elastic range. Such dissipation is due to various mechanisms
such as cracking, nonlinearity in the elastic phase of response, interaction with
non-structural elements, soil-structure interaction, etc. As it is very difficult and
unpractical to estimate each mechanism individually, the elastic viscous damping
represents the combined effect of all of the dissipation mechanisms. There is no
direct relationship of such damping with the real physical phenomena. However, the
adoption of the viscous damping concept facilitates the solution of the differential
equations of motion, represented by Eq. (1)

ü + 2ξn u̇ + n2 u = 0, (1a)

where
c
ξ= , (1b)
2mn

u is the displacement, ωn is the natural vibration frequency (radians/sec) of the


system and ξ is the damping ratio or fraction of critical damping.
As mentioned above, the assumption of viscous damping simplifies greatly the
dynamic problem and this is the reason why in the direct displacement based design
the nonlinear behaviour has been represented by including the hysteretic damping
into the damping term of Eq. (1a). Using this approach it is possible to solve a
simple linear system instead of a nonlinear system which is more time and resource
demanding for design applications.

2.1. Modelling of hysteretic damping with equivalent


viscous damping
An early proposal to model the inelastic behaviour with a parameter proportional
to the velocity was made by Jacobsen [1930, 1960]. He approximated the nonlinear
friction behaviour to a power of velocity. This was initially used to compute the
response of single-degree-of-freedom (SDOF) systems when subjected to sinusoidal
loads. Housner [1956] and Jennings [1964] carried out investigations in order to
apply the concept to other hysteretic systems. As a general start point, the total
equivalent viscous damping equations that have been proposed by other authors
Equivalent Viscous Damping Equations 259
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

(a) (b)
Fig. 1. Dissipated and stored force for (a) viscous damping and (b) hysteretic cycles.

(see Sec. 2.3) are divided in two parts (Eq. (2)):


ξeq = ξo + ξhyst , (2)
where ξo corresponds to the initial damping in the elastic range and ξhyst cor-
responds to the equivalent viscous damping ratio that represents the dissipation
due to the nonlinear (hysteretic) behaviour. In this study, the first variable will be
ignored as explained subsequently.
For the equivalent viscous damping corresponding to the hysteretic response, the
concept of dissipated (EDiss ) and stored (Esto ) energy has been used [Jacobsen,
1930]. With reference to Fig. 1, the value of the equivalent viscous damping ratio
can be obtained equating the energy dissipated by a viscous damper with the energy
dissipated from nonlinear behaviour:
1 n EDiss
ξhyst = · . (3)
4π  Esto
In order to use this approach it is necessary to assume that both systems are
subjected to harmonic excitation. This is necessary to ensure that the loops are
complete and to obtain a closed-form solution for the displacement. Additionally,
it is assumed that the excitation frequency is the same as the natural frequency of
the SDOF system (resonance condition). The final result is thus given by Eq. (4).
1 EDiss 1 Ahyst
ξhyst = · = . (4)
4π Esto 2π Fo uo
It is clear, however, that response to real earthquake excitation cannot be exactly
represented by steady-state harmonic response, and that an unknown error will be
introduced in the estimation of displacements, based on the approximations made
in Jacobsen’s approach.

2.2. Previous research


Gulkan and Sozen [1974] extended Jacobsen’s approach by introducing the concept
of substitutive viscous damping (SVD) based on limited experimental results. They
260 C. A. Blandon & M. J. N. Priestley

obtained the value of this factor for a given secant stiffness (effective period) and a
given time history by balancing the input energy of the SDOF with a linear dashpot
that would bring the system to rest (Eq. (5)):
t
Tsubstitute 0 üg · u̇ · dt
ξsubstitute = t , (5)
4π 0 u̇2 dt
where u is the structure displacement, t is the total time of the accelerograms and
ü g is the ground acceleration. Tsubstitute is the effective period corresponding to
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

the secant stiffness to maximum response and ξsubstitute is the SVD. They also
computed the same factor using the approximation by Jacobsen and found that
the results were not significantly different. More extensive analyses by Judi et al.
[2000] found significant differences between Jacobsen’s Equivalent viscous damping
factor, and Gulkan and Sozen’s substitute damping factor, and concluded that
designs based on the latter provided a better estimate of the expected displacement
response.
Kowalsky and Ayers [2002] also investigated the substitute damping approach,
and concluded that it was necessary to carry out additional investigation in order
to find out the limits and variation of using this simplified assumption. They also
carried out research about the equivalent damping of concrete structures, attempt-
ing to identify potential limitations and the range of applicability of Jacobsen’s
equivalent damping for DDBD. Based on the initial stages of this research they
found that “on average, assessment of nonlinear response with equivalent linear
systems (defined by effective period at maximum response and equivalent damp-
ing defined by Jacobsen’s approach) yields good results for the majority of cases
considered”. However they found that in cases with acceleration records with large
velocity pulses, the equivalent damping approach failed to recognize that the peak
nonlinear response is no longer a function of the energy dissipated. This effect was
also pointed out by Priestley [2003] when he proposed a modified reduction equation
for high damping levels for the spectral displacements for earthquakes containing
large velocity pulses.
In a more recent research, Kowalsky and Dwairi [2004] concluded that the
Jacobsen approach frequently overestimated the damping and that the fundamen-
tal period of the system, the characteristics of the ground motion and the ductility
level are critical variables for the equivalent damping concept.

2.3. Existing equivalent viscous damping equations


There are multiple references which report different equations for the equivalent vis-
cous damping factor, including those by Priestley [2003], Fardis and Panagiotakos
[1996], Miranda and Ruiz [2002], Calvi [1999]. Some are based on Jacobsen’s
approach, some on the substitute damping approach, and others on results of time-
history analyses. Some of the reported equations are listed below and compared
in Fig. 2.
Equivalent Viscous Damping Equations 261

50 Rosenblueth

Equivalent Damping Factor


45
Steel*
40
35 Conc. beam *
30
Conc. Wall*
25
20 Kow alsky
15 Gulkan
10
5 Iw an
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

0 Prestr . Conc.*
1 2 3 4 5 6 7
Ductility

Fig. 2. Equivalent viscous damping factor for Eqs. (6a) to (6h). Proposed by Priestley [2003]
r = 0.05 for those equations requiring it.

Bilinear elasto-plastic system, Rosenblueth and Herrera [1964]:


 
2 (1 − r)(μ − 1)
ξeq = ξo + . (6a)
π μ − rμ + rμ2
Takeda model, Gulkan Sozen [1974]:
 
1
ξeq = ξo + 0.2 1 − √ . (6b)
μ
Elastic and Coulomb slip elements, Iwan [1980]:
ξeq = ξo + 0.0587(μ − 1)0.371 . (6c)
Takeda model α = 0.5 and β = 0, Kowalsky [1994]:
 
1 1−r √
ξeq = ξo + 1− √ −r μ . (6d)
π μ
Steel members, Priestley [2003]:
150
ξeq = 5 + (μ − 1). (6e)
μπ
Concrete frames, Priestley [2003]:
 
120 1
ξeq =5+ 1− √ . (6f)
π μ
Concrete columns, and walls, Priestley [2003]:
 
95 1
ξeq = 5 + 1− √ . (6g)
π μ
Precast walls or frames, with unbonded prestressing: Priestley [2003]:
 
25 1
ξeq = 5 + 1− √ . (6h)
π μ
262 C. A. Blandon & M. J. N. Priestley

In general, the equations proposed by Priestley have the form (Eq. (7)):
 
1
ξeq = ξo + a 1 − β , (7)
μ

where r is the post yielding stiffness coefficient, ξeq is the equivalent viscous damping
factor, ξo is the initial viscous damping and μ is the ductility level.
Miranda [2002] carried out an investigation comparing the capabilities of differ-
ent performance based methodologies (including methodologies based on equivalent
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

linearization) in estimating the inelastic displacement for Takeda, modified Clough,


stiffness degrading and elastoplastic system using 264 ground motion records. The
findings of Miranda also suggested that Jacobsen’s approach for estimating the
equivalent viscous damping factor was nonconservative for structures with high hys-
teretic energy absorption. It is important to point out that the equivalent damping
from Eqs. (6) and (7) are related to specific definition of the structural period.
Almost all the expressions given in Eq. (6) are defined for secant stiffness at maxi-
mum response, but the period related to the expression (6c) is given by an empirical
equation that matches the results of a set of nonlinear dynamic analyses carried
out by Iwan; Additional details can be found elsewhere [Blandon, 2004; Miranda,
2002; Kowalsky, 2002].
The studies mentioned above consistently indicated problems with use of
Jacobsen’s approach for equivalent viscous damping within DDBD methodology.
It is then necessary to carry out additional research in order to complement previ-
ous studies and to optimise damping values applicable to a wide range of hysteretic
models and period ranges. This study attempts such an optimization.

3. Accelerograms and Displacement Spectra


Six different synthetic accelerograms were selected in order to carry out the time-
history analysis of the SDOF systems. All of them were constructed matching a
given displacement design response spectrum at 5% damping. The first accelero-
gram was a synthetic record based on a record from Manjil, Iran, 20th June 1990,
developed by Bommer and Mendis [2004] for use specifically in this research. This
spectrum is shown in Fig. 3(a) for different levels of damping. The rest of the
accelerograms were artificial records obtained as part of this study and comple-
mented with others obtained by Alvarez [2004] and Sullivan [2003]. The average
response spectra for these five records are shown in Fig. 3(b) for different lev-
els of damping. In these figures only 6 levels of damping are shown; however, for
the analyses, the average spectra for 20 levels of damping were computed for val-
ues between 3% and 60%. The spectral value for a given damping was obtained
by interpolation between the closest upper and lower spectra bounding it. All of
the records were compatible with the ATC32 design spectrum for soil type C
and PGA 0.7 g. [ATC32, 1996].
Equivalent Viscous Damping Equations 263

1-Spectral Analysis
3- Dynamic Analysis
Accelerograms
Accelerograms

Equivalent Select μ
Damping, ξ eq
Select Teff Hysteresis Model
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

Damped
Displacement Max Displacement
Spectra
Dhyst

Average Spectral
Displacement (Dspec)

4- Convergence Check
2- System Design
Dhyst= Dspec
SDOF Systems
Ky, Fy yes
No

Keep Equivalent Change Equivalent


Go to 3 Damping Damping

Fig. 3. Flux diagram for the proposed methodology.

The spectrum-compatible records used for the study were selected carefully so
that the results of the dynamic analyses are representative of a larger set of non-
spectrum compatible records.

4. Hysteretic Models
Six hysteretic models were considered in the analysis: Elastic perfectly plastic
(EPP), bilinear type model, a “narrow” and a “fat” Takeda loop, a Ramberg Osgood
model and finally a ring spring (flag shape) model (Fig. 4).
The equivalent viscous damping equations based on Jacobsen’s approach for five
of these models are given in Eqs. (8a) to (8d).
Modified Takeda Model [Loeding et al. (1998)]:
       
2 3 1 rβμ 1 1
ξequ = 1 − μα−1 − 1− +1 2−β· 1− − μα−1 γ
π 4 4 γ μ μ
  2
1 rβ 2 μ 1 (8a)
− 1− ,
4 γ μ

γ = rμ − r + 1,
264 C. A. Blandon & M. J. N. Priestley

5%
Displacement

Δ spec
Reduced
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

Teff

Period

Fig. 4. Selection of the displacement from the averaged reduced spectrum.

where r is the post yielding stiffness ratio, μ is the displacement ductility, α is


the unloading stiffness parameter and β is the reloading stiffness parameter. Two
versions were used in the analyses. The first, a “thin” Takeda model with α = 0.5
and β = 0 was considered appropriate for bridge piers and wall structures. The
second “fat” Takeda model (α = 0.3, β = 0.6) was intended to be appropriate for
reinforced concrete frames. In both cases the post-yield stiffness ratio was taken
as 0.05.
Bilinear [Otani, 1981] r:
2(μ − 1)(1 − r)
ξequ = . (8b)
πμ(1 + r(μ − 1))
The Bilinear model was intended to model a bridge structure isolated with
friction pendulum dampers. The behaviour represented the composite flexibility
of dampers plus bridge piers, and a high post-yield stiffness ratio of r = 0.2 was
consequently adopted.
Elastic-Perfectly-Plastic (EPP) [Otani, 1981]:
2(μ − 1)
ξequ = . (8c)
π·μ
The EPP model was included solely because of its historic importance in seis-
mic time-history analysis. Its closest approximation in real structures is a flexible
structure isolated with a flat coulomb (friction) damper.
Ramberg Osgood [Otani, 1981]:
   
1 2 1 F
ξequ = 1− · 1− · , (8d)
π γ+1 μ Fy
where γ is the Ramberg Osgood Parameter [see Fig. 4(e)]. This was selected as
being reasonably appropriate for structural steel members.
Equivalent Viscous Damping Equations 265

Ring Spring:
The Ring-spring model [Fig. 6(f)] was chosen to represent a precast concrete struc-
ture connected with unbonded prestressing, resulting in a hysteretic model char-
acterised by bilinear elastic response with low hysteretic damping. The equation
corresponding to the ring spring model is more complex. It was obtained with the
use of a mathematical software and compared with that provided by Kowalsky and
Ayers [2002] and is not shown here because of space limitation, but is available in
Blandon [2004].
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

5. Methodology for Calibration of Equivalent Viscous Damping


for DDBD
Based on the results obtained from the research described in Sec. 2, a methodology
was implemented, to modify, where necessary, Jacobsen’s equations for the equiv-
alent viscous damping factor (ξ). The procedure adopted determined the value of
equivalent damping that has to be applied to an equivalent elastic system with
a given effective period (based on the secant stiffness to maximum displacement
response) in order to match its response (in terms of maximum displacement) to
that obtained from a system with the same period (effective period) and a given
level of ductility using nonlinear time-history analysis. The final objective of this
procedure was to develop equations that define the equivalent damping factor to be
used in DDBD for a given level of ductility and hysteretic model. The flow diagram
of the proposed method is shown in Fig. 5.

Step 1: Initially, an effective period (Teff ) and a ductility level (μ) are selected.
Step 2: Estimate the equivalent damping factor ξ. For the first iteration this was
based on Jacobsen’s approach according to the hysteretic loop considered. How-
ever, after the results of the first iteration were obtained, the equivalent damping

Manjil Average Response Spectra


1.2 1.2

1 1.0
5
DIsplacement (mm)
Displacement (m)

0.8 0.8 10
15

0.6 0.6 20
30

0.4 0.4 40
Caltrans 5%

0.2 0.2

0 0.0
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(a) (b)
Fig. 5. Displacement response spectra adopted for analysis. (a) Modified Manjil Record [Bommer
and Mendis, 2004]. (b) ATC32 [1996] (Averaged).
266 C. A. Blandon & M. J. N. Priestley

was changed in the next iterations to improve the substitute-structure/time-history


agreement.
Step 3: Determine the average damped displacement spectrum for the calculated
value of ξ based on the average spectra obtained from the accelerograms described
in Sec. 4.
Step 4: An initial response displacement (Δspec ) is obtained from the actual average
damped spectrum (non-smoothed) for the selected effective period Teff as shown
in Fig. 4.
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

Step 5: For a given hysteretic model, the initial stiffness (Kini ) and yielding force
(Fy ) are defined using Δspec , mass (meff ), effective period (Teff ) and the ductility
(μ) as follows:
Yielding displacement (Eq. (9)):

Δspec
Δy = . (9)
μ

Secant stiffness (Eq. (10)):

Keff = 4π 2 meff /Teff


2
. (10)

Maximum Force (Eq. (11)):

Fmax = Keff Δspec . (11)

The yield force Fy can be found from the ductility and maximum response force
using hysteretic-model-specific equations. These equations are given in Sec. 5.4. The
initial stiffness Kini can then be found from:
Fy
Kini = . (12)
ΔY
The mass used in this step was used only so that correct ductility was obtained. In
the analyses the mass was taken constant, but there is not a conceptual issue by
selecting this parameter.
Step 6: Run time-history analysis for each of the records and obtain the maximum
displacements.
Step 7: Compare the displacements obtained from Step 6 with that from Step 4.
Step 8: If the displacements are similar (within a tolerance of 3%), keep the damping
factor used and repeat the process from Step 1 with the next Teff and μ, otherwise,
modify the damping factor and repeat the process from Step 2.
The process is repeated for effective periods from 0.5 s to 4 s each 0.5 s, for
5 ductility levels from 2 to 6. Six different hysteretic curves are used and all the
cases are analyzed for six records (Fig. 6).
Equivalent Viscous Damping Equations 267

F dp dp
Fy F β dp
rKo
Fy

Ku
Ko
Ko Ku=Ko (dy /dm)α
D
dy dm D
Ku=Ko (dy /dm)α dy dm
Ku
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

rKo - Fy rK - Fy
o

(a) (b)

Fy
F

rK0 Ko Ko

Q
K0 D
Keff
D

- Fy

(c) (d)

F/Fy F
⎛ Do Fo ⎛
⎜ , ⎜
λ −1
⎜D F⎜
D
=
F
(1 + η) F ⎝ y y⎝
D y Fy Fy
dy d0
D/Dy D
⎛ Do Fo⎛ rlower K0
⎜− ,− ⎜ F0
⎜ D Fy⎜⎝
⎝ y
λ −1 rsteep K0
D − Do F − Fo F − Fo
= (1 + η) Fy
2D y 2 Fy 2 Fy rK0

(e) (f)
Fig. 6. Hysteretic models. (a) Thin Takeda (r = 0.05, α = 0.5 and β = 0.0). (b) Fat Takeda
(r = 0.05, α = 0.3 and β = 0.6). (c) Bilinear (r = 0.2). (d) EPP (r = 0). (e) Ramberg Osgood
(γ = 7). (f) Ring spring (rlower = 0.035, rsteep = 1, r = 0.04).
268 C. A. Blandon & M. J. N. Priestley

6. Modelling and Analysis Assumptions


The equivalent viscous damping in DDBD is defined as the combination of two
effects as described in Eq. (2). There is the elastic damping and the hysteretic
damping.
The initial elastic viscous damping used for time-history analysis of SDOF sys-
tems has been traditionally defined in practice by use of a constant damping coef-
ficient corresponding to 5% of critical damping, though lower values are sometimes
used for steel structures. This value is assumed to represent the different sources
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

of energy dissipation when the structure is considered in the elastic range. It is


not clear that constant coefficient damping is appropriate for structures respond-
ing inelastically, since the hysteretic models generally incorporate the full structural
energy dissipation in the inelastic range, and other contributory mechanism, such as
foundation damping will be greatly reduced when the structure enters the inelas-
tic range. It would appear that tangent-stiffness proportional damping would be
more appropriate than constant coefficient (initial-stiffness proportional, or mass-
proportional) damping in modelling initial elastic damping in seismic response. The
adoption of different characteristic stiffness in DDBD (secant stiffness) and time-
history analyses (initial stiffness) further confuses the issue.
The representation of elastic damping is often given little consideration, but as
shown in the accompanying paper by Priestley and Grant [2005], the issue of how
this damping should be represented is important, not just to DDBD, but also to
time-history analysis. In this study the influence of elastic damping was removed
from both the design process and the time-history validation, by specifying zero
elastic damping.
Time-history analysis were carried out using the program RUAUMOKO [Carr,
2003], using a Newmark constant average acceleration integration scheme with
β = 0.25. As described in Sec. 5, this procedure was carried out iteratively until
the displacement of the equivalent SDOF system was the same for the time-
history and for the design spectral analysis. The time step used for the integra-
tion was taken as half of the discretization step of each accelerograms, this is,
0.005 seconds except for the Manjil adjusted record which has a discretization
step of 0.0045.

7. Results
The results shown in this section are the average of the individual analysis for
each of the six accelerograms used for each hysteretic model. Two set of plots are
analysed for each case; the first set presents the ratio between the displacements
obtained from the time-history analysis (THA) and those from the initial step of
the spectral DDBD procedure basing the equivalent viscous damping on Jacobsen’s
approach). From this ratio it is possible to estimate the expected error in the DDBD
when based on Jacobsen’s approach. The second set of plots show the effective
Equivalent Viscous Damping Equations 269

equivalent viscous damping factor, obtained from the iterative procedure, necessary
to equate the design displacement and the average results from the time-history
analyses.

7.1. Time history displacements versus design displacements


using Jacobsen’s approach
Results for the six hysteretic models investigated are presented in Fig. 7, where
the ratio is plotted against effective period at maximum displacement for different
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

ductility levels. It is seen that the ratio of average time-history displacement over
design displacement based on Jacobsen’s approach depends on hysteresis model,
ductility level, and period. The agreement is better for the models with low hys-
teretic energy (Thin Takeda, Fig. 7(a), and ring-spring, Fig. 7(f)) For these, the
average ratio is generally within ±10% of unity. As a general rule the ratio increases
with area in the hysteretic loop, and with period. The variation with ductility
level is less clear, with some models (e.g. Ramberg Osgood, Fig. 7(e)) showing an
increase in the ratio with ductility level, and others (e.g. Bilinear, Fig. 7(c)) show-
ing a reduction. For the models with high energy dissipation, particularly the EPP
model (Fig. 7(d)) the average displacement ratios are excessive, and exceed 2.0 in
some cases. It should be noted that these ratios are expected to be exaggerated by
the removal of elastic damping from the design and time-history analyses.
The C.O.V. between the different records was highly dependent on the hysteretic
rule. For Takeda and Bilinear model it was between 10% an 20% on average, while
for the Ramberg Osgood and the ring spring it was between 20 to 25%. Finally, for
the EPP model it increased to between 40 to 50% approximately, indicating the
unreliability of this hysteretic model.

7.2. Effective damping factor to equate design and


time-history displacements
The effective damping factor necessary to equate design and time-history displace-
ments, obtained from the iterative procedure is plotted against period, for different
ductilities in Fig. 8, and against ductility for different periods, in Fig. 9. Also shown
in Fig. 9 are the relationships given by Jacobsen’s equations, and where appropri-
ate, by Priestley’s equations (Eqs. 6(e) to 6(h)). The results indicate an increase
with ductility in all cases, and a general tendency for the damping to reduce with
increasing period. The existing equations are reasonable for models with low hys-
teretic energy absorption, but overestimate the damping by as much as 100% for the
models with high hysteretic energy absorption. Note that there are significant differ-
ences between the equivalent viscous damping levels for different hysteretic models
for the same ductility level. This would not be the case if the “equal-displacement”
approximation of structural response was valid.
For the equal displacement to be consistent with the results the equivalent
damping should be very similar for a given period and a given ductility level in any
270 C. A. Blandon & M. J. N. Priestley

1.4 1.4

1.3 1.3

Displacement Ratio
2 1.2
Displacement Ratio

1.2

THA/Jacobsen
THA/Jacobsen

3 3
1.1 1.1 4
4
5 5
1.0 1.0
6 6

0.9 0.9
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

0.8 0.8
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(a) (b)

1.5
4.0
1.4
3.5
2
1.3
Displacement Ratio

3 3.0
THA/Jacobsen

2
Displacement Ratio

1.2 4 2.5
THA/Jacobsen

3
1.1 5 2.0 4
6 1.5 5
1.0
6
1.0
0.9
0.5
0.8
0.0
0 1 2 3 4 5
0 1 2 3 4 5
Period (s) Period (s)

(c) (d)

2.2 1.3

2.0
1.2
1.8
2 2
Displacement Ratio

Displacement Ratio
THA/Jacobsen

THA/Jacobsen

3 1.1
1.6 3
4 4
1.4
5 1.0 5
1.2 6 6
0.9
1.0

0.8 0.8
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(e) (f)
∗ Series represent the ductility level
Fig. 7. Time history analysis/Initial design displacement average ratio (a) Takeda model (Narrow
type, α = 0.5, β = 0.0, r = 0.05), (b) Takeda model (Fat type, α = 0.3 and β = 0.6, r = 0.05),
(c) Biliniear (r = 0.2), (d) EPP (r = 0), (e) Ramberg Osgood (γ = 7), (f) Ring spring, based on
Jacobsen’s Approach. Series represent the ductility level.
Equivalent Viscous Damping Equations 271

hysteretic model. This occurs just for some periods and ductility level but there
can be large differences as in the case of the ring-spring model.

8. Design Equations for Equivalent Viscous Damping Factor


From the results presented in the previous section, it is clear that it is necessary
to develop revised expressions for the equivalent viscous damping to be used in
DDBD. Although it would be possible to use the results from Figs. 8 and 9 directly,
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

it is clear that this would be a cumbersome approach, and design equations would
be preferable. It is also clear that simple modification of the equations representing
Jacobsen’s approach (Eq. (8)) would not be practical because of the complexity of
some of the equations, and the lack of period-dependency.
The equations proposed by Priestley were taken as the base for the modified
equations proposed here. The resulting equations exist in the following form:
1
ξeffective = f (μ) · f (T ) · , (13a)
N
   
a 1 1 1
ξeffective = · 1− b · 1+ · , (13b)
π μ (T + c)d N
where a, b, c and d are coefficients defined for each hysteretic model, μ is the
ductility, T is the effective period and N is a normalising factor. An important
difference of this equation from previous existing proposals is the extra term which
is dependant on effective period. An exception exists in the recommendations of
Judi et al. [2002], which indicate weak period-dependency.
The function f (μ) was matched as closely as possible to the values of the effective
damping for an effective period of 0.5 s. This function was then modified by the f (T )
in order to match the damping to the other periods. The normalizing factor (N ) is
just the result of the expression f (T ), evaluated for T = 0.5:
1
N = 1+ . (14)
(0.5 + c)d
It is important to mention that the implementation of this approach in DDBD will
modify slightly the design process. This is because, in the methodology presented
in the companion paper by Pettinga and Priestley [2005], the damping is obtained
directly from the ductility; then, the effective period is obtained for a given target
displacement. However, using the modified equation it will be necessary to iterate
in order to obtain the period. The additional work is, however, insignificant.
The process in obtaining the calibration factors a, b, c, d was carried out for
each hysteretic model analysed. Not all the correction factors depend, in the same
proportion, on the variables (period and ductility). A perfect match was not possible
for all the cases because it was necessary to keep a simple form of the equation.
Further in the case of the bilinear model, it was necessary to modify the basic form
of Eq. (13) to obtain a match of adequate accuracy.
272 C. A. Blandon & M. J. N. Priestley

20 30
18
25
16
Damping Factor %

Damping Factor %
14 2 2
20
3 3
12
4 4
10 5 15 5
8 6 6
10
6
4
5
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

2
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(a) (b)

25 35

30
20
Damping Factor %

Damping Factor %

25 2
2 3
15 3 20 4
4 5
10 5 15 6
6
10
5
5

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(c) (d)

30 10
9
25
8
Damping Factor %

Damping Factor %

2 7 2
20
3 3
6
4 4
15 5 5 5
6 4 6
10
3
2
5
1
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(e) (f)
∗ Series represent the ductility level
Fig. 8. Average equivalent damping versus period for (a) Takeda model (Narrow type, α = 0.5,
β = 0.0, r = 0.05), (b) Takeda model (Fat type, α = 0.3 and β = 0.6, r = 0.05), (c) Bilinear
(r = 0.2), (d) EPP (r = 0), (e) Ramberg Osgood (γ = 7), (f) Ring spring.
Equivalent Viscous Damping Equations 273

20 30

18 0.5 25
1.0 0.5

Damping Factor (%)


Damping Factor (%)

16
1.5 1.0
2.0 20 1.5
14 2.0
2.5
2.5
12 3.0 15 3.0
3.5 3.5
10 4.0 4.0
JB 10 JB
8 Priest.
Priest.
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

6 5
2 3 4 5 6 2 3 4 5 6
Ductility Ductility

(a) (b)

25 55
50
20 45
0.5 0.5
Damping Factor (%)
Damping Factor (%)

40
1.0 1.0
1.5 35 1.5
15
2.0 30 2.0
2.5 25 2.5
10 3.0 20 3.0
3.5 15 3.5
4.0 4.0
5 10
JB JB
5
0 0
2 3 4 5 6 2 3 4 5 6
Ductility Ductility

(c) (d)

40 10

35 9
8
0.5 0.5
Damping Factor (%)

30
Damping Factor (%)

7
1.0 1.0
25 1.5 6 1.5
2.0 2.0
20 5
2.5 2.5
15 3.0 4 3.0
3.5 3 3.5
10 4.0 4.0
JB 2 JB
5 Priest. 1 Priest.

0 0
2 3 4 5 6 2 3 4 5 6
Ductility Ductility

(e) (f)
∗ Series represents the effective period
Fig. 9. Average Equivalent damping versus ductility for (a) Takeda model (Narrow type, α = 0.5,
β = 0.0, r = 0.05), (b) Takeda model (Fat type, α = 0.3 and β = 0.6, r = 0.05), (c) Bilinear
(r = 0.2), (d) EPP (r = 0), (e) Ramberg Osgood (γ = 7), (f) Ring spring.
274 C. A. Blandon & M. J. N. Priestley

Table 1. Constant values for Eq. (13) or Eq. (15) for each hysteretic model.

Takeda Takeda Ramberg Ring


Constant Thin Fat Bilinear EPP Osgood Spring
a 95 130 160 140 150 50
b 0.5 0.5 0.5 0.5 0.45 0.5
c 0.85 0.85 0.85 0.85 1 1
d 4 4 4 2 4 3
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

The effective damping obtained for the bilinear model (Sec. 7) show some char-
acteristics which are somewhat different from the other models. The relationship
between the EVDF and the ductility tends to have a different shape for ductility
levels larger than 4. In order to follow this behaviour, it was necessary to adjust
the general equation (Eq. (13)) as follows:
   
a 1 1 1
ξeffective = · 1 − b − 0.1 · r · μ · 1 + · , (15)
π μ (T + c)d N

where r is the post elastic stiffness coefficient. However, it would be necessary to


carry out additional analyses in order to determine for which range of values of
r this equation is valid. Table 1 shows the constants obtained for the hysteretic
models analysed.
The agreement between damping levels predicted using Eqs. (13) or (15) and
the results of the iterative procedure are shown in Fig. 10. Values predicted by the
equations are shown by dashed lines with the time-history results by solid lines. It
will be noted that the agreement is satisfactory over the full range of periods and
ductility levels, for each hysteretic model.
One alternative to this procedure could be derived from the fact that the depen-
dency of the equivalent damping reduces very rapidly as the period increases. The
damping becomes almost independent for a period larger than 1 sec. Using this
approximation it would not be necessary to iterate in the design procedure as
mentioned previously. However, given that the procedure is simple the proposed
equation has the advantage of being general for any period. Reduced version of the
equation may be obtained easily according to the characteristics of the structures
considered in the design.

9. Conclusions
The design displacements based on Jacobsen’s approach within the direct
displacement-based design methodology were calculated for six different hysteretic
models with different effective periods and displacement ductilities. These design
displacements were compared with results from time-history analyses using six
Equivalent Viscous Damping Equations 275

20 30

2 25 2
Damping Factor %

Damping Factor %
15
3 20 3

4 4
10 15
5 5
10
5 6 6
5

0 0
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(a) (b)

30 35

25 2 30 2
Damping Factor %

Damping Factor %

3 25 3
20
4 20 4
15
5
15 5
10
10
6 6
5 5
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(c) (d)

30 10

25 2 8 2
Damping Factor %

Damping Factor %

20 3 3
6
4 4
15
5 4 5
10
6 6
5 2

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Period (s) Period (s)

(e) (f)
∗ Series
represent the ductility level (Dashed lines represent the damping factor estimated by the
modified equation for each ductility level and effective period)
Fig. 10. Modified EVDF for (a) Takeda model (Narrow type, α = 0.5, β = 0.0, r = 0.05),
(b) Takeda model (Fat type, α = 0.3 and β = 0.6, r = 0.05) (c) Bilinear (r = 0.2), (d) EPP (r = 0),
(e) Ramberg Osgood (γ = 7), (f) Ring spring.
276 C. A. Blandon & M. J. N. Priestley

earthquake records compatible with the design spectra. Both designs and time-
history analyses were carried out using zero elastic viscous damping to enable the
contribution of hysteretic damping to be directly determined.
As has been found in earlier studies, the results obtained were inconsistent, with
the time-history displacements exceeding the design displacements in many cases,
particularly for hysteretic models with high energy absorption.
An iterative procedure was used to determine the required value for equivalent
viscous damping to be used in direct displacement-based design to equate the design
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

displacement and the time-history results. From these analyses a series of design
equations determining the equivalent viscous damping as a function of hysteresis
rule, displacement ductility and period were developed.
It was found that the equal displacement approximation is not always consistent
with the results obtained. Damping for different hysteretic models shows that the
use of this approximation can be unreliable.
Final verification of the approach when the response is combined with elas-
tic viscous damping is on-going. This requires consideration of the form of elastic
damping most appropriate to model real structural behaviour, and a recognition
that the formulation of direct displacement-based design uses a different character-
istic stiffness than that in time-history analysis. This influences the numeric value
of elastic damping to be added to the hysteretic damping to model (say) 5% elastic
damping in the real structure.

Acknowledgements
The authors are grateful to Professor Dr. Julian Bommer and Ing. Rishmilla Mendis
from Imperial College for the hard work carried out in order to generate some of
the records used for the analysis.

References
Abrahamson, N. A. [1998] Non-Stationary Spectral Matching Program RSPMATCH,
PG&E Internal Report, February.
Alvarez, J. C. and Priestley, M. J. N. [1994] Displacement-Based Design of Continuous
Concrete Bridges under Transverse Seismic Excitation, MSc Thesis, ROSE School,
Pavia Italy.
ATC32 [1996] Applied Technology Council, Improved Seismic Design Criteria for
California Bridges: Provisional Recommendations, Rept. No. ATC-32, Redwood City,
California.
Blandon, C. A. [2004] Equivalent Viscous Damping for DDBD, MSc Thesis, ROSE School,
Pavia Italy.
Bommer, J. J. and Mendis, R. [2004] “Scaling of displacement spectral ordinates with
damping ratios,” Earthquake Engineering & Structural Dynamics (in press).
Borzi, B. et al. [2001] “Inelastic spectra for displacement based seismic design,” Soil
Dynamics and Earthquake Engineering 21, 47–61.
Equivalent Viscous Damping Equations 277

Calvi, G. M. [1999] “A displacement-based approach for vulnerability evaluation of classes


of buildings,” Journal of Earthquake Engineering 3(3), 411–438.
Carr, A. J. [2002] SIMQKE Generator of Artificial Earthquakes, University of Canterbury,
Christchurch, New Zealand.
Chopra, A. K. [1995] Dynamics of Structures: Theory and Application to Earthquake Engi-
neering, Prentice Hall, USA.
EuroCode 8 [1988] Structures in Seismic Regions — Design, Part 1, General and Building,
May 1988 Edition, Report EUR 8849 EN, Commission of the European Communities.
Fardis M. N. and Panagiotakos, T. B. [1996] “Hysteretic damping of reinforced concrete
elements,” Elsevier Science Ltd, 1996, Eleventh World Conference on Earthquake
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

Engineering, Paper No. 464.


Gulkan, P. and Sozen M. A. [1974] “Inelastic responses of reinforced concrete structures
to earthquakes motions,” Proceedings of the ACI 71 (12).
Housner, G. W. [1956] “Limit design of structures to resist earthquakes,” Proceedings of
the First World Conference on Earthquake Engineering, San Francisco 5, 1–13.
Iwan, W. D. and Gates, N. C. [1979] “The effective period and damping of a class
of hysteretic structures,” Earthquake Engineering and Structural Dynamics V(7),
199–211.
Jacobsen, L. S. [1930] “Steady forced vibration as influenced by damping,” Transactions
of ASME 52, 169–181.
Jacobsen, L. S. [1960] “Damping in composite structures,” Vol 2, Proceedings of the Second
World Conference on Earthquake Engineering, pp. 1029–1044.
Jennings, P. C. [1964] “Periodic response of a general yielding structure,” Journal of the
Engineering Mechanics Division, ASCE 90(2), 131–166.
Judi, H. J., Davidson, B. J. and Fenwick, R. C. [2000] “The direct displacement based
design — A damping perspective,” 12th World Conference on Earthquake Engineer-
ing, Auckland, New Zealand, January 2000, Paper No. 0330.
Judi, H., Fenwick, R. C. and Davidson, B. J. [2002] “Influence of hysteretic form on seismic
behaviour of structures,” Proceeding on NZSEE Conference, Napier, paper No. 6.5,
p. 10.
Kowalsky, M. and Dwairi, H. [2004] “Investigation of Jacobsen’s equivalent viscous damp-
ing approach as applied to displacement-based seismic design,” 13th World Conference
on Earthquake Engineering, Vancouver, Canada, Paper 228.
Kowalsky, M. J. and Ayers, J. P. [2002] “Investigation of equivalent viscous damping for
direct displacement-based design,” PEER-2002/02, The Third US-Japan Workshop
on Performance-Based Earthquake Engineering Methodology for Reinforced Concrete
Building Structures, 16–18 August 2001, Seattle, Washington, Berkeley: Pacific Earth-
quake Engineering Research Center, University of California, pp. 173–185.
Loeding, S., Kowalsky, M. J. and Priestley, M. J. N. [1998] Direct Displacement-Based
Design of Reinforced Concrete Building Frames, University of California, San Diego,
La Jolla, CA.
Research Project SSRP-98/08, 297 Pages.
Miranda, E. and Ruiz-Garcia, J. [2002] “Evaluation of the approximate methods to
estimate maximum inelastic displacement demands,” Earthquake Engineering
and Structural Dynamics 31, 539–560.
Otani, S. [2002] Nonlinear Analysis of Reinforced Concrete Buildings (Lecture notes), Rose
School, Pavia, Italy.
Otani, S. [1981] “Hysteresis models of reinforced concrete for earthquake response analy-
sis,” Journal of Faculty of Engineering, University of Tokyo XXXVI(2), 407–441.
278 C. A. Blandon & M. J. N. Priestley

Pettinga, D. and Priestley, M. J. N. [2005] “Dynamic behaviour of reinforced concrete


frames designed with direct displacement-based design,” Journal of Earthquake Engi-
neering, Special Edition.
Priestley, M. J. N. and Calvi, G. M. [1997] “Concepts and procedures for direct
displacement-based design,” Seismic Design Methodologies for the Next Generation
of Codes, Fajfar and Krawinkler (eds.) Balkema, Rotterdam, 171–181.
Priestley, M. J. N. [2003] “Myths and fallacies in earthquake engineering, revisited,” The
Mallet Milne Lecture. IUSS Press, Pavia, Italy.
Priestley, M. J. N. and Grant, D. N. [2005] “Viscous damping for analysis and design,”
Journal of Earthquake Engineering, Special Edition.
Downloaded by [University of Saskatchewan Library] at 11:54 23 August 2013

Shibata, A. and Sozen, M. A. [1976] “Substitute-structure method for seismic design in


R/C,” Journal of the Strutural Division, ASCE, New York, USA.
Structural Engineers Assn. of California (SEAOC), Vision 2000 Committee [1995] “Perfor-
mance based seismic engineering of buildings,” J. Soulages (ed). 2 vols. [Sacramento,
Califorina.]
Sullivan, T. [2003] “The limitations and performances of different displacement based
design methods,” Journal of Earthquake Engineering, 7 (Special Issue 1), 201–241.

You might also like