Amplitude-Dependent Damping in Vibration Serviceability: Case of A Laboratory Footbridge

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/295920414

Amplitude-Dependent Damping in Vibration Serviceability: Case of a Laboratory


Footbridge

Article  in  Journal of Architectural Engineering · February 2016


DOI: 10.1061/(ASCE)AE.1943-5568.0000211

CITATIONS READS
22 816

1 author:

Onur Avci
Iowa State University
85 PUBLICATIONS   1,932 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Dynamics, Railway, Vibrations View project

Metamaterial Inspired Structures (Metastructures) for Vibration Suppression View project

All content following this page was uploaded by Onur Avci on 12 December 2017.

The user has requested enhancement of the downloaded file.


Amplitude-Dependent Damping in Vibration Serviceability:
Case of a Laboratory Footbridge
Onur Avci, Ph.D., P.E., M.ASCE1

Abstract: Construction technology advancements in the last couple of decades have led to the use of lightweight and high-strength
materials in structural systems. Although longer spans and lighter materials result in floor systems with less mass, stiffness, and damp-
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

ing, the trend toward a paperless office decreases damping and the amount of live load on the floors even more. Consequently, structures
have become more vulnerable to annoying vibrations, and vibration serviceability has become an area of serviceability concern. For
vibration serviceability calculations, the damping value of the structural systems is a critical parameter. Damping in structures has
proved to be dependent on the amplitude of the applied force on the structure. This condition is referred to as nonlinear damping, or
amplitude-dependent damping. Although damping is constant at low and high amplitudes, for in-between amplitudes, the damping value
increases with the levels of excitation amplitude. For wind and earthquake excitations, the amplitude-dependent characteristics of
damping have been studied extensively in the literature. For floor vibration serviceability applications, even though the nonlinear behav-
ior of damping has been accepted to exist and mentioned in some publications, it is not closely looked at or discussed in detail. The floor
vibration serviceability calculations are very sensitive to damping values, but vibration serviceability researchers and practicing engi-
neers are often uncomfortable with assigning a specific number as a damping ratio for a specific mode because of the inconsistency of
damping values obtained from different methods. This paper presents a closer look at the amplitude-dependent damping in vibration
serviceability and focuses on a laboratory footbridge with experimental and analytical studies. The laboratory footbridge was studied
extensively with static and dynamic tests. Three-dimensional finite-element (FE) models were developed, updated, and fine-tuned for
two bottom chord extension configurations for both static and dynamic tests. The amplitude-dependent damping behavior of the labora-
tory footbridge is shown for different amplitudes of sinusoidal excitations. The amplitude-dependent damping ratio values obtained
from effective mass calculations proved to be correct with the FE model acceleration predictions. The FE model predictions successfully
matched the test results with the nonlinear characteristic introduced for modal damping. One of the most difficult tasks in vibration serv-
iceability research is matching the measured acceleration responses with the FE models, and the success of this paper in matching the
acceleration responses for various levels of excitations (with corresponding amplitude-dependent damping values) with the FE model is
unique. Successful verification and clarification of the amplitude-dependent phenomenon and FE model matching of measured accelera-
tion responses reinforce the confidence in the FE models in vibration serviceability research by showing that the FE models are reliable
not only for natural frequency predictions but also for acceleration response predictions. DOI: 10.1061/(ASCE)AE.1943-
5568.0000211. © 2016 American Society of Civil Engineers.

Introduction values would underestimate the true acceleration response of a


structure, which might result in problems after the design stage.
Technological advancements in building construction have facili- Generally, damping can be classified as material and nonmate-
tated the use of lightweight and high-strength materials in structural rial damping (Sun and Lu 1995). Material damping is caused by
systems over the years. This resulted in larger bays, longer spans, the energy dissipation property of the material via hysteresis loops,
and more slender structures with less damping ratios. In addition, during cyclic deformation. For instance, high-damping alloys,
with more demanding architectural design trends, the gravity sys- fiber-reinforced composite materials, and viscoelastic materials are
tem of structures has become more vulnerable to annoying vibra- the type of materials with large hysteresis loops. For typical conven-
tions. For more than a couple of decades, annoying vibrations have tional structural components, however, the hysteresis loop is very thin
been an area of serviceability concern in structural design. within the elastic range of the materials. Also, it must be noted that
For vibration serviceability calculations, the damping value of material damping is independent of the excitation frequency.
the gravity system is a key parameter. However, the current model- Nonmaterial damping is a type of damping caused by the sur-
ing techniques and descriptions are unsuccessful at predicting rounding medium in which the vibration takes place. Coulomb
damping. Dealing with damping has always been very critical for damping, viscous damping, and radiation damping are common
researchers and practitioners because overestimating the damping types of nonmaterial damping. Coulomb damping (dry damping) is
caused by the relative motions of connected components at the con-
1
tact surfaces (e.g., structural joints). Viscous damping, which is des-
Assistant Professor, Dept. of Civil and Architectural Engineering,
ignated by a dashpot in structural dynamics applications, is caused
Qatar Univ., P.O. Box 2713, Doha, Qatar. E-mail: onur.avci@qu.edu.qa
Note. This manuscript was submitted on May 11, 2015; approved on
by the dissipation of energy that occurs when a vibration movement
January 11, 2016; published online on February 26, 2016. Discussion pe- is resisted by a force with a magnitude proportional to the magni-
riod open until July 26, 2016; separate discussions must be submitted for tude of the velocity of the system and acts in an opposite direction
individual papers. This paper is part of the Journal of Architectural to the velocity direction. The radiation damping is the energy dissi-
Engineering, © ASCE, ISSN 1076-0431. pation that occurs at the supports and connections of the structure; it

© ASCE 04016005-1 J. Archit. Eng.

J. Archit. Eng., 04016005


is basically the mechanism that propagates energy away from the modes in it, whereas the modal damping ratio from frequency
structure (Pavic 1999). The vibration movement of the structure response functions (FRFs) contains only a single mode. Yet, this
applies dynamic forces at the support locations, which causes the ra- does not explain the disagreement within the frequency domain
diation of stress waves into the support and beyond. While prediction methods. It is also agreed that widely and closely spaced
Coulomb damping and radiation damping are independent of exci- modes show different relationships between modal damping val-
tation frequency, viscous damping is dependent on the excitation ues obtained from time and frequency domains (Rainer and Van
frequency (Ewins 2000). Selst 1976). Although floor vibration calculations are extremely
Typically, because the decay curves of real structures are very sensitive to modal damping values, researchers and practitioners in
similar to the viscous model curves, for the sake of simplicity, a this field are not very comfortable with assigning one specific num-
structure’s overall damping is assumed to be like the viscous ber to be a damping ratio for a specific mode. While Pavic (1999)
damping model. Wyatt (1977) suggested that there are many fric- noted that it is hard to predict damping, Tilly et al. (1984) and
tional damping mechanisms in structures, and the sum of all damp- Živanovic et al. (2005) suggested specifying the measured damp-
ing mechanisms conform to the viscous damping model. In prac- ing values with the corresponding acceleration response ampli-
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

tice, Wyatt’s model shows that stuck-friction models predict an tudes. Even though this was a very good idea, it has not been put in
amplitude-dependent damping characteristic. practice in floor vibration serviceability applications.
Damping has been known to be dependent on the amplitude of the This paper presents a closer look at amplitude-dependent damp-
applied force on the structure. De Silva (2005) defined amplitude- ing and focuses on the nonlinear behavior of damping by studying a
dependent damping as a “nonlinear form of damping where its laboratory footbridge without any nonstructural components on it.
value depends on the amplitude but not on the frequency of The laboratory footbridge was studied extensively with static and
motion.” Although damping is constant at low and high amplitudes, dynamics tests. Tests were conducted without any nonstructural
for in-between amplitudes, the damping value increases with the ex- components attached to focus only on the amplitude-dependent
citation amplitude. This nonlinear behavior of damping has been damping behavior of the bare structure. Three-dimensional finite-
studied by many researchers, such as Leonard and Eyre (1975), element (FE) models were developed, updated, and fine-tuned for
Jeary and Ellis (1981, 1983), Jeary (1986, 1996, 1997), Chang and different bottom chord extension configurations for both static and
Mohraz (1990), Zhang and Zhang (1994), Bachmann et al. (1995), dynamic tests. The amplitude-dependent characteristic of the foot-
Morita and Kanda (1996), Tamura and Suganuma (1996), Li et al. bridge is shown for different amplitudes of sinusoidal excitations.
(2000, 2003), Butterworth et al. (2004), Živanovic et al. (2005), Initially, three-dimensional FE models were developed and
Brownjohn and Pavic (2007), Wu et al. (2007), Racic et al. (2009), updated for each bottom chord extension configuration to reflect the
Daoulatli (2010), Díaz and Reynolds (2010a, b), Middleton and static test results of the following:
Brownjohn (2010), Ingólfsson et al. (2011), Jang (2011, 2013), • Midspan point loading of bare joists (deflections and bottom
Hudson and Reynolds (2012), Sharma et al. (2012), Taillon et al. chord extension forces are matched with the FE model);
(2012), Aquino and Tamura (2013), Casini et al. (2013), Gromysz • Wet concrete loading on bare joists (deflections and bottom
(2013), Spence et al (2014), Zapico-Valle et al. (2013), Ho et al. chord extension forces are matched with the FE model);
(2014), Steinwolf et al. (2014), Zoghaib and Mattei (2014), and • Midspan point loading on cured concrete slab (deflections and
Salyards and Noss (2014). bottom chord extension forces are matched with the FE
According to Jeary (1996), damping is a result of the mobiliza- model); and
tion of imperfections within the vibrating structure. From this per- • Uniformly distributed loading on cured concrete slab (deflec-
spective, damping will increase linearly with the vibration ampli- tions and bottom chord extension forces are matched with the
tude but it will remain constant at very low and very high vibration FE model).
amplitudes. When a structure is excited, initially, the damping will After the static testing, three-dimensional FE models were
be constant until the large imperfections are mobilized (with low updated for each bottom chord extension configuration to reflect the
amplitudes of vibration). As the vibration amplitude increases, dynamic test results of the following:
damping will also increase as relatively smaller imperfections • Natural frequency of governing flexural bending mode.
within the structure are mobilized. With additional increase in the Being in excellent correlation with the previous test series, the
vibration amplitude, the smallest imperfections in the system will FE models were fine-tuned and detailed enough to reflect the
be mobilized and the damping will be constant for further increases changes in the footbridge structural system due to removal of
in the excitation amplitude. Even though some structures seem to bottom chord extensions from the footbridge structure. The results
conform to Jeary’s damping model (Jeary 1996), it is very difficult of the static testing are already shared by Avci and Murray (2012).
to quantify the damping values accurately (Middleton and The partial results of the dynamic testing (modal parameter estima-
Brownjohn 2010). Similarly, Smith (2001) noted the nonlinearity of tions) are also shared by Avci (2014).
damping and mentioned that “it is often difficult to obtain a precise The study presented in this paper focuses on the amplitude-
value for damping.” dependent damping characteristics of the laboratory footbridge.
For wind and earthquake excitations on structures, the amplitude- After the peak picking method, the natural frequency for the gov-
dependent characteristics of damping have been studied exten- erning bending mode had been determined; however, the nonli-
sively in the literature. However, for floor vibration serviceability nearity in modal damping values was unexplained. The damping
applications, even though the nonlinear behavior of damping has value discrepancies between the vibration software curve fits
been accepted and mentioned in some publications, it is not closely (MEScope) (Vibrant Technology, Inc. 2003), individual FRF
looked at or discussed in detail. The nonlinear damping behavior half-power method results, and time domain decay curve values
has been observed, and the phenomenon has been simplified in are explained and verified in this paper. The footbridge was put
such a way that sometimes the damping values from the frequency in resonance by applying sinusoidal excitations at the governing
domain (half-power method and curve fitting) and time domain bending mode frequency. From the effective mass calculations
(decay curves) do not perfectly match. A very simple explanation of the resonance conditions, the nonlinear behavior characteris-
for this is that the decay curve includes the response from several tics of the damping are identified and verified with the FE model

© ASCE 04016005-2 J. Archit. Eng.

J. Archit. Eng., 04016005


Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Laboratory footbridge, 2.13  9.14 m (7  30 ft)


Fig. 4. Modal testing setup (images by Onur Avci)

Fig. 5. Accelerometer locations

Fig. 2. Bottom chord extension locations on the footbridge

Fig. 6. First bending mode shape (FE model)


(a)

by showing that they are reliable not only for natural frequency
predictions but also for acceleration response predictions in vibra-
(b) tion serviceability research.

Fig. 3. Bottom chord extension configurations for the footbridge:


(a) Stage 1, bottom chord extensions in place; (b) Stage 2, bottom chord Footbridge Test Setup
extensions removed
The footbridge (Fig. 1) was constructed at the Structural Engi-
neering Laboratory, Virginia Tech, Blacksburg, Virginia. The
acceleration predictions. The FE model predictions successfully dimensions of the footbridge are 2.13  9.14 m (7  30 ft). The
matched the test results with the nonlinear characteristic intro- cold-formed steel deck is a 1.5VL-type deck (Vulcraft, Norfolk,
duced for the modal damping ratio. Nebraska) with a depth of 38 mm (1.5 in.) and a normal weight
While one of the difficult tasks in floor vibration serviceability concrete slab with a depth of 114 mm (4.5 in.), with a total slab
research is matching the measured acceleration responses with the depth of 152 mm (6.0 in.) supported on two parallel lines of
FE models, the success of this paper in matching the acceleration 30K7  9.14-m (30-ft) span Vulcraft joists at 1.22 m (4 ft) on cen-
responses for various levels of excitations (with corresponding ter. Grade 8 stand-off screws (Elco Construction Products,
amplitude-dependent damping values) with the FE model is Decorah, Iowa) with 102 mm (4 in.) of stand-off and spaced at
unique. Successful verification and clarification of the amplitude- 305 mm (12 in.) on center were used to connect the cold-formed
dependent phenomenon and FE model matching of measured steel decks to the joist top chords before the concrete was placed.
acceleration responses reinforces the confidence in the FE models The 28-day concrete compressive strength is 29.8 MPa (4,320

© ASCE 04016005-3 J. Archit. Eng.

J. Archit. Eng., 04016005


1
2
3

4
5

6
7
8 10
9 11 13
12 14
15

Fig. 7. First bending mode shape (MEScope model)


Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Stage 1 FRF for center point chirp excitation and center point accelerometer

Fig. 9. Stage 2 FRF for center point chirp excitation and center point accelerometer

© ASCE 04016005-4 J. Archit. Eng.

J. Archit. Eng., 04016005


psi). Reinforced concrete walls were used as supports. The joist response predictions. When the FE model predictions match the test
seats were welded to the bearing plates located on the concrete results, the FE model is verified and can be used with more confi-
walls at the supports. Bottom chord extension members were dence (Ewins 2000; Inman 2000). The goal of the modal testing
HSS 38  38  4.8 mm (1.5  1.5  3/16 in.). The bottom chord conducted in this study was to determine the modal parameters,
extensions were instrumented as load cells so that the axial force such as natural frequencies, modal damping ratios, and mode shapes
in the member could be monitored, recorded, and matched in FE of the laboratory footbridge, and to determine the effect of bottom
models. Two of the four bottom chord extensions of the foot- chord extensions for the two configurations. For that purpose, chirp
bridge (S1 and S2 in Fig. 2) were instrumented with load cells. excitations were applied on the footbridge to find the governing
In the experimental program, static and dynamic tests were con- bending mode frequencies for each bottom chord extension stage.
ducted on the footbridge. For the simulation of static and dynamic The footbridge was excited with a shaker at a location, and the
tests, three-dimensional FE models of the footbridge were created acceleration response was measured at various locations. The FRFs
using the SAP2000 commercial software program (CSI 2002). are computed by using the acceleration response and input force. In
The flexural stiffness tests were conducted first to understand this study, the input force excitation is measured by a force plate.
the flexural stiffness behavior of the footbridge and then to start
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

The shaker was placed on the force plate to measure the force
tuning the FE models. The footbridge was tested with two differ- applied on the footbridge (Fig. 4). The data from modal testing were
ent bottom chord extension configurations (stages) shown in then loaded, processed, and curve fitted in MEScope commercial
Fig. 3. The two stages of the footbridge are the bottom chord software (Vibrant Technology, Inc. 2003) for both stages, and natu-
extension configurations needed to determine the relative effect ral frequencies, modal damping ratios, and the corresponding mode
of installing bottom chord extensions after the concrete is placed.
shapes were approximated per the input data. MEScope is commer-
Static flexural stiffness testing protocol, FE modeling techni-
cial software used for observing and analyzing noise and vibration
ques, and stiffness test results are shown in detail by Avci and
problems in machinery and structures using either experimental or
Murray (2012), therefore, they will not be repeated in this manu-
analytical data. It can import or directly acquire multichannel time
script. The static stiffness test results (with and without concrete
or frequency data from a machine or a structure and postprocess it
slab) were verified with the FE models for the bottom chord
to come up with operating deflection shapes, resonant vibration,
extension configurations.
and mode shapes from acquired data.
For modal testing, the shaker and force plate assembly were
Modal Testing of the Footbridge Structure placed at different locations on the footbridge, and chirp signals
were used to sweep a frequency range of 4–20 Hz for the two bot-
The aim of modal testing is to apply dynamic excitations to the tom chord extension configurations. The first bending frequency
structure and update the FE model based on the test results with the was the main interest because it is the only mode that falls in the
intent of minimizing the discrepancies between them for improved range that governs the acceleration response of the structure for

Table 1. First Bending Mode Natural Frequencies and Damping Ratios for the Footbridge

First bending mode frequency, fn (Hz) Modal damping ratio ( b )


Curve fitting of Half-power Curve fitting of Time domain decay curves
Individual vibration data FE model method vibration data (filtered to capture first
Stage FRF (MEScope) (SAP2000) (individual FRF) (MEScope) mode only)
1 8.00 8.08 7.99 0.00284 0.00451 0.01300
2 6.95 6.95 7.03 0.00453 0.00448 0.02200

20

0
dB = 20*log10(abs(XferDat))

-20

-40

-60 Magnitude 2/1(V/V)


Magnitude 3/1(V/V)
Magnitude 4/1(V/V)
Magnitude 5/1(V/V)
-80 Magnitude 6/1(V/V)
Magnitude 7/1(V/V)
Magnitude 8/1(V/V)
-100
0 2 4 6 8 10 12 14 16 18 20
Frequency (Hz)

Fig. 10. Chirp FRF for Stage 1 in logarithmic scale

© ASCE 04016005-5 J. Archit. Eng.

J. Archit. Eng., 04016005


human perception. Fig. 5 shows the locations on which the accel- are shown in Figs. 6 and 7, respectively. For Stage 1, the FRF
eration response data were collected. The first bending mode plots for the center point chirp excitation and the center point ac-
shapes with the FE model and the curve-fitted MEScope model celerometer location are shown in Fig. 8. The same plot for Stage

5.4
4.9988
5.0 Xfer Actual

Transfer Function Magnitude


4.6
4.2
4.9988 / 21/2
4.9988 / 21/2
3.8 =3.5347
=3.5347
3.4
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

3.0
2.6
Half Power Method:
2.2 β = [8.125- 8.079] / (2x8.10)
β = 0.284 %
1.8
1.4
8.05 8.06 8.07 8.08 8.09 8.10 8.11 8.12 8.13 8.14 8.15
Frequency (Hz)

Fig. 11. FRF for Stage 1: zoomed in for half-power method calculations

0.045
0.035
Filtered:
0.025 above 10 Hz
below 6 Hz
Acceleration (g)

0.015
0.005
-0.005
-0.015 Curve Fit
β = 1.3%
-0.025 Filtered

-0.035
-0.045
1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time (s)

Fig. 12. Stage 1 heel drop excitation decay curve and the fitted curve

ap
p0
System Response
Input Excitation

Time Time

Fig. 13. Input excitation at resonant frequency and the corresponding system response for damped systems

© ASCE 04016005-6 J. Archit. Eng.

J. Archit. Eng., 04016005


2 is shown in Fig. 9. The magnitude and real and imaginary parts bottom chord extensions from the system (Stage 2) resulted in a
of the FRF are shown as well as the phase and coherence informa- 14% drop (from 8.08 to 6.95 Hz) in the natural frequency of the
tion in Figs. 8 and 9. Reporting the phase and coherence informa- footbridge. Stage 1 natural frequency is higher than that of Stage 2,
tion is important in such measurements to show the quality of the which means the bottom chord extensions do stiffen the footbridge.
FRFs (Avitabile 2008). The FE model natural frequency predictions are also shown in
MEScope approximations for the natural frequency and modal Table 1. It is noted that the curve-fitted MEScope and FE model
damping ratios are shown in Table 1. It is shown that removing the (SAP2000) frequency predictions are in excellent agreement (within

Force Plate Time History and Autospectrum Time History and Autospectrum of Accelerometer Data
30 0.2

Accelerometers (g)
Force Plate (lbs)

0.1
20 Sinusoidal excitation
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

amplitude Sinusoidal acceleration response amplitude = 0.14 g


0
= 12.7 lbs
10
−0.1

0 −0.2
0 5 10 15 20 0 5 10 15 20
Time (sec) Time (sec)

0.08 0.2
Autospectrum FP

0.06 (Peak)x(1.414)= Time Domain Force Amplitude 0.15 AUTOSPECTRUM

Autospectrum
Channel 1 x 1.414 x (224.3 lb/Volt)= 12.4 lbs Peak Values: (Peak)x(1.414)=Time Domain Amplitude
Peak Ch 2 = 0.099425 (Peak Ch 2 x 1.414) = 0.14059 g
0.04 0.1 Peak Ch 3 = 0.016365 (Peak Ch 3 x 1.414) = 0.02314 g
X: 8
Y: 0.03902
Peak Ch 4 = 0.019753 (Peak Ch 4 x 1.414) = 0.02793 g
Peak Ch 5 = 0.098821 (Peak Ch 5 x 1.414) = 0.13973 g
0.02 0.05 Peak Ch 6 = 0.10041 (Peak Ch 6 x 1.414) = 0.14198 g
Peak Ch 7 = 0.040583 (Peak Ch 7 x 1.414) = 0.057384 g
Peak Ch 8 = 0.04239 (Peak Ch 8 x 1.414) = 0.05994 g
0 0
0 5 10 15 20 0 5 10 15 20
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 14. Sinusoidal excitation at 8.0 Hz, amplitude = 56.5 N (12.7 lbs), acceleration response amplitude = 0.14 g: (a) input; (b) output

Force Plate Time History and Autospectrum Time History and Autospectrum of Accelerometer Data
40 0.4
Accelerometers (g)
Force Plate (lbs)

0.2
20 Sinusoidal excitation amplitude Sinusoidal
= 26.7 lbs acceleration
0 response
amplitude
0 = 0.26 g
−0.2

−20 −0.4
0 5 10 15 20 0 5 10 15 20
Time (sec) Time (sec)

0.1 0.2

0.08
Autospectrum FP

(Peak)x(1.414)= Time Domain 0.15 AUTOSPECTRUM


X: 8
Autospectrum

Y: 0.08371

Peak Values: (Peak)x(1.414)=Time Domain Amplitude


Force Amplitude (Peak Ch 2 x 1.414) = 0.25374 g
0.06 Peak Ch 2 = 0.17945
Channel 1 x 1.414 x (224.3 lb/Volt) (Peak Ch 3 x 1.414) = 0.043983 g
0.1 Peak Ch 3 = 0.031105
= 26.6 lbs Peak Ch 4 = 0.035527 (Peak Ch 4 x 1.414) = 0.050235 g
0.04 Peak Ch 5 = 0.18 (Peak Ch 5 x 1.414) = 0.25451 g
Peak Ch 6 = 0.17983 (Peak Ch 6 x 1.414) = 0.25428 g
0.05 (Peak Ch 7 x 1.414) = 0.1076 g
0.02 Peak Ch 7 = 0.076099
Peak Ch 8 = 0.083473 (Peak Ch 8 x 1.414) = 0.11803 g
0 0
0 5 10 15 20 0 5 10 15 20
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 15. Sinusoidal excitation at 8.0 Hz, amplitude = 118.8 N (26.7 lbs), acceleration response amplitude = 0.26 g: (a) input; (b) output

© ASCE 04016005-7 J. Archit. Eng.

J. Archit. Eng., 04016005


Force Plate Time History and Autospectrum
Time History and Autospectrum of Accelerometer Data
50
0.4

Accelerometers (g)
Force Plate (lbs)
Maximum Absolute Force: 0.2 Sinusoidal
Channel 1 = 44.7 lbs
acceleration
0 0 response amplitude
= 0.36 g

−0.2

−50 −0.4
0 5 10 15 20 0 5 10 15 20
Time (sec) Time (sec)
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

0.2 0.4
(Peak)x(1.414)=Time Domain Amplit.
(Peak Ch 2 x 1.414) = 0.35861 g
Autospectrum FP

(Peak)x(1.414)= Time Domain AUTOSPECTRUM


0.15 0.3 Peak Values: (Peak Ch 3 x 1.414) = 0.063455 g

Autospectrum
Force Amplitude
(Peak Ch 4 x 1.414) = 0.072783 g
Channel 1 x 1.414 x (224.3 lb/Volt)= Peak Ch 2 = 0.25361
(Peak Ch 5 x 1.414) = 0.35976 g
0.1 44.1 lbs 0.2 Peak Ch 3 = 0.044877 (Peak Ch 6 x 1.414) = 0.35911 g
Peak Ch 4 = 0.051473
(Peak Ch 7 x 1.414) = 0.16109 g
Peak Ch 5 = 0.25443
(Peak Ch 8 x 1.414) = 0.17858 g
0.05 0.1 Peak Ch 6 = 0.25397
Peak Ch 7 = 0.11393
Peak Ch 8 = 0.1263
0 0
0 5 10 15 20 0 5 10 15 20
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 16. Sinusoidal excitation at 8.0 Hz, amplitude = 198.8 N (44.7 lbs), acceleration response amplitude = 0.36 g: (a) input; (b) output

Table 2. Stage 1: Sinusoidal Excitations Test Data Sinusoidal Excitations for Resonance
Sine excitation amplitude
Acceleration response amplitude Neither the frequency domain nor the time domain methods pro-
N lbs (test) (%g) vided a consistent damping ratio value for the footbridge. The
56.5 12.7 0.14 damping ratios from the half-power method, MEScope curve fits,
118.8 26.7 0.26 and decay curves do not provide clear information. The most reli-
198.8 44.7 0.36 able method for achieving more consistent modal damping ratios
remains the effective mass calculations when the footbridge
structure is put in resonance. This can be done by applying excita-
1.5% of each other for Stages 1 and 2). It is also observed from tions at the structure’s natural frequency. Therefore, sinusoidal
Table 1 that the frequency domain methods for damping ratio pre- excitations were applied to the footbridge through a shaker in an
dictions (half-power method and MEScope curve-fitted damping attempt to put the structure in resonance. The resonance condition
ratios) do not consistently match each other. The time domain would allow the use of the effective mass procedure to calculate
method of the decay curve yields a much higher damping ratio than modal damping ratios. A sinusoidal excitation input will result in
the frequency domain predictions for the two stages (even though a sinusoidal acceleration response on the footbridge at the reso-
the decay curve was filtered to contain the first bending mode con- nant frequency (Fig. 13).
tent only).
For the FE model acceleration predictions, damping ratio values Resonant Condition for Stage 1
need to be entered for the footbridge models. However, based on
The footbridge was sinusoidally excited with three different lev-
the scatter observed in Table 1, it is not clear which damping ratio
els of excitation amplitudes (56.5, 118.8, and 198.8 N) for this
should be used in the FE models. The damping ratio for the first
stage. The shaker was placed at the center of the footbridge, and
bending mode cannot be confidently used with the inconsistency
sine waves with an 8.0-Hz frequency were applied to the struc-
shown in Table 1.
ture. The footbridge was put in resonance. The sinusoidal excita-
The FRF in the logarithmic plot is shown in Fig. 10. Fig. 11
tion time history and acceleration responses were measured. The
shows the zoomed-in view for the peak of the FRF plot for half- force amplitudes and the resulting acceleration responses are
power method calculations. Fig. 12 shows the time domain heel shown in Figs. 14–16. The resonant test results for Stage 1 are
drop decay curve for a heel drop conducted at the center of the struc- tabulated in Table 2.
ture. The curve fit to the decay curve yields a damping ratio of
1.3%.
To clarify the modal damping ratio inconsistency, it is decided Resonant Condition for Stage 2
to apply sinusoidal excitations to the footbridge and put the struc- The bottom chord extensions were removed from the system
ture in resonance. When the footbridge is in resonance, the effective and dynamic testing continued. The footbridge was sinusoi-
mass procedure can be used to get the modal damping ratios. The dally excited with four different levels of excitation ampli-
excitation force amplitudes and the resulting acceleration ampli- tudes (33.8, 53.4, 120.1, and 195.7 N) for this stage. The
tudes will be used for the effective mass calculations. shaker was placed at the center of the footbridge, and sine

© ASCE 04016005-8 J. Archit. Eng.

J. Archit. Eng., 04016005


Force Plate Time History and Autospectrum
Time History and Autospectrum of Accelerometer Data
Force Plate (lbs) 70 Sinusoidal 0.2
excitation

Accelerometers (g)
65 Sinusoidal
amplitude= 0.1 acceleration
7.6 lbs response
60 0 amplitude=
0.12g
55
−0.1
50
−0.2
5 10 15 5 10 15
Time (sec) Time (sec)
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

0.2 0.1
Autospectrum FP

0.15 (Peak)x(1.414)= Time Domain Force Amplitude 0.08

Autospectrum
AUTOSPECTRUM (Peak)x(1.414)=Time Domain Amplit.
Channel 1 x 1.414 x (224.3 lb/Volt)= 7.59lbs Peak Values: (Peak Ch 2 x 1.414) = 0.11958 g
0.06 Peak Ch 2 = 0.084572 (Peak Ch 3 x 1.414) = 0.013773 g
0.1 Peak Ch 3 = 0.0097404 (Peak Ch 4 x 1.414) = 0.013526 g
Peak Ch 4 = 0.0095656 (Peak Ch 5 x 1.414) = 0.12287 g
0.04 Peak Ch 5 = 0.086896 (Peak Ch 6 x 1.414) = 0.12006 g
0.05 X: 6.95
Y: 0.02393
Peak Ch 6 = 0.08491 (Peak Ch 7 x 1.414) = 0.015746 g
0.02 Peak Ch 7 = 0.011136 (Peak Ch 8 x 1.414) = 0.019461 g
Peak Ch 8 = 0.013763
0
0
5 10 15 0 5 10 15 20
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 17. Sinusoidal excitation at 6.95 Hz, amplitude = 33.8 N (7.6 lbs), acceleration response amplitude = 0.12 g: (a) input; (b) output

Force Plate Time History and Autospectrum Time History and Autospectrum of Accelerometer Data
80
Accelerometers (g)

0.2
Force Plate (lbs)

70 Sinosoidal Sinusoidal
excitation 0.1 acceleration
60 amplitude = response
12 lbs 0
50 amplitude
−0.1 = 0.17g
40
−0.2
30
5 10 15 5 10 15
Time (sec) Time (sec)

0.2
0.15 Maximum Absolute Force:
AUTOSPECTRUM
Autospectrum FP

Channel 1 = 32.8173 lbs


Autospectrum

Time Domain force amplitude 0.15 Peak Values: (Peak)x(1.414)=Time Domain Amplitude
= 0.03904 x 224.3 x 1.414 = Peak Ch 2 = 0.11958 (Peak Ch 2 x 1.414) = 0.16909 g
0.1 Peak Ch 3 = 0.01376 (Peak Ch 3 x 1.414) = 0.019456 g
12.4 lbs X: 6.95
Y: 0.1196(Peak Ch 4 x 1.414) = 0.018602 g
0.1 Peak Ch 4 = 0.013156
(Peak Ch 5 x 1.414) = 0.1739 g
Peak Ch 5 = 0.12298
0.05 Peak Ch 6 = 0.1201 (Peak Ch 6 x 1.414) = 0.16982 g
X: 6.95
0.05 Peak Ch 7 = 0.016026 (Peak Ch 7 x 1.414) = 0.02266 g
Y: 0.03904
Peak Ch 8 = 0.019963 (Peak Ch 8 x 1.414) = 0.028227 g
0
0
5 10 15 0 5 10 15 20
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 18. Sinusoidal excitation at 6.95 Hz, amplitude = 53.4 N (12.0 lbs), acceleration response amplitude = 0.17 g: (a) input; (b) output

waves with a 6.95-Hz frequency were applied to the structure. Effective Mass Calculations for Resonant Conditions
The footbridge was put in resonance. The sinusoidal excitation
time history and acceleration responses were measured. The When a system is vibrating at a natural frequency, its deforma-
force amplitudes and the resulting acceleration responses are tion follows a specific pattern, which is called a mode shape.
shown in Figs. 17–20. The resonant test information for Stage For each mode shape the amount of mass contributing to the
2 is tabulated in Table 3. system response is different. Effective mass is an indication of

© ASCE 04016005-9 J. Archit. Eng.

J. Archit. Eng., 04016005


Force Plate Time History and Autospectrum Time History and Autospectrum of Accelerometer Data
100

Accelerometers (g)
Force Plate (lbs)
0.2 Sinusoidal
80 Sinusoidal acceleration
excitation response
60 amplitude 0 amplitude =
= 0.24g
40 27.0 lbs −0.2

20
0 5 10 15 20 0 5 10 15
Time (sec) Time (sec)
0.4 0.2
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

Autospectrum FP

AUTOSPECTRUM

Autospectrum
X: 6.95
0.3 0.15 Peak Values:
Y: 0.1725
(Peak)x(1.414)=Time Domain Amplitude
(Peak)x(1.414)= Time Domain Force AmplitudeMaximum Absolute Force:
Channel 1 = 28.7393 lbs Peak Ch 2 = 0.17246 (Peak Ch 2 x 1.414) = 0.24386 g
Channel 1 x 1.414 x (224.3 lb/Volt)= 27.3 lbs Peak Ch 3 = 0.0203 (Peak Ch 3 x 1.414) = 0.028704 g
0.2 0.1 Peak Ch 4 = 0.019474 (Peak Ch 4 x 1.414) = 0.027536 g
Peak Ch 5 = 0.12298 (Peak Ch 5 x 1.414) = 0.1739 g
Peak Ch 6 = 0.1201 (Peak Ch 6 x 1.414) = 0.16982 g
0.1 0.05 Peak Ch 7 = 0.016026 (Peak Ch 7 x 1.414) = 0.02266 g
X: 6.95 Peak Ch 8 = 0.019963 (Peak Ch 8 x 1.414) = 0.028227 g
Y: 0.0862

0 0
0 5 10 15 20 0 5 10 15 20
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 19. Sinusoidal excitation at 6.95 Hz, amplitude = 120.1 N (27.0 lbs), acceleration response amplitude = 0.24 g: (a) input (b) output

Force Plate Time History and Autospectrum Time History and Autospectrum of Accelerometer Data
0.4
Accelerometers (g)

80
Force Plate (lbs)

Sinusoidal 0.2 Sinusoidal


60 excitation acceleration
amplitude 0 response
40
= 44.0 lbs amplitude =
20 −0.2 0.32g

0
−0.4
0 5 10 15 20 0 5 10 15 20
Time (sec) Time (sec)
0.2 0.4
Autospectrum FP

(Peak)x(1.414)=Time Domain Amplitude


Sinusoidal excitation force AUTOSPECTRUM
Autospectrum

0.15 (Peak Ch 2 x 1.414) = 0.32494 g


0.3 Peak Values:
X: 6.95 amplitude = Peak Ch 2 = 0.2298
(Peak Ch 3 x 1.414) = 0.038835 g
Y: 0.14 (Peak Ch 4 x 1.414) = 0.03853 g
0.14 x 224.3 x 1.414 = 44 lbs Peak Ch 3 = 0.027465
(Peak Ch 5 x 1.414) = 0.12287 g
0.1 0.2 Peak Ch 4 = 0.027249 X: 6.95
Y: 0.2298
(Peak Ch 6 x 1.414) = 0.12006 g
Peak Ch 5 = 0.086896
(Peak Ch 7 x 1.414) = 0.015746 g
Peak Ch 6 = 0.08491
(Peak Ch 8 x 1.414) = 0.019461 g
0.05 0.1 Peak Ch 7 = 0.011136
Peak Ch 8 = 0.013763

0 0
0 5 10 15 20 0 5 10 15 20
(a) Frequency (Hz) (b) Frequency (Hz)

Fig. 20. Sinusoidal excitation at 6.95 Hz, amplitude = 195.7 N (44.0 lbs), acceleration response amplitude = 0.32 g: (a) input; (b) output

mass sensitivity of each mode to an excitation. Mode shapes Table 3. Stage 2: Sinusoidal Excitation Test Data
can be calculated experimentally using modal data. Each ac-
celerometer location is assigned a lumped mass of the tested Sine excitation amplitude
Acceleration response amplitude
structure (Fig. 21), and collected chirp data for each lumped N lbs (test) (%g)
mass location can then be loaded into commercial vibration
33.8 7.6 0.12
analysis software (MEScope) and the mode shapes can be
53.4 12.0 0.17
determined (Fig. 22).
120.1 27.0 0.24
When a mode shape of a structure is known, the effec-
195.7 44.0 0.32
tive mass corresponding to that mode can be calculated.

© ASCE 04016005-10 J. Archit. Eng.

J. Archit. Eng., 04016005


2 3
m=8 0 0 0 0
6 7
6 0 m=4 0 0 7
0
6 7
M¼6
6 0 0 m=4 0 7
0
7
6 7
4 0 0 0 m=4 0 5
0 0 0 0 m=8 (4)

where m = total mass of the footbridge.


For the footbridge, acceleration data were collected on three
parallel lines in the longitudinal direction. For the effective mass
calculations, the average mode shape of the three lines is used and
the system is treated as a beam. The mode shape vectors for the
first bending mode of the footbridge were determined for both
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

stages. Each mode shape vector was normalized to establish the


f matrix of Eq. (3), and the effective mass of the entire foot-
bridge for the first bending mode was calculated. The results are
shown in Table 4.
From the MEScope mode shapes, Stage 1 resulted in an effective
mass of 0.470 m for the first bending mode of the footbridge.
Removing the bottom chord extensions from the system (Stage 2)
resulted in a 7% increase (from 0.470 to 0.502 m) in the effective
mass for the first bending mode, as shown in Table 4. (This was
expected because Stage 2 is stiffer.)
As mentioned earlier, when a structure is excited with a har-
monic driving force at its natural frequency, the response will be
resonant. It is known that for a system excited with a harmonic force
at its natural frequency, the resulting peak acceleration is
p0
ap ¼ (5)
2 b Meff

Fig. 21. Mode shape determination where ap = peak acceleration; p0 = excitation amplitude of the sinu-
soidal forcing function; b = modal damping ratio; and Meff = effec-
tive mass for a mode.
Effective mass of a mode in a two-way structural system is Eq. (5) can also be written as the following:
defined as ap p0
¼ (6)
g 2 b Weff
Lðx Lðy

Meff ¼ Mðx; yÞð f ðx; yÞÞ2 dxdy (1) where Weff = the effective weight for a mode.
0 0 The total weight of the footbridge structure is 6,344 kg
(13,986 lbs).
Considering Eq. (6), the ap and p0 values are known from the
Effective mass of a mode in a one-way structural system (like
sinusoidal excitation test results for Stages 1 and 2 (Tables 2 and
the footbridge discussed in this paper) is defined as
3). When the effective weight values of 0.470 and 0.502 W are
used for Stages 1 and 2, respectively, the corresponding damping
ðL
ratios are found for each sinusoidal test run, using Eq. (6). Table
Meff ¼ MðxÞð f ðxÞÞ2 dx (2) 5 shows the calculated damping ratios for Stage 1, and Table 6
0 shows the calculated damping ratios for Stage 2. It is very clear
from Tables 5 and 6 that as the sine excitation amplitude
where MðxÞ = mass per unit length of the system (a constant for increases, the damping ratio value also increases. This is the am-
floor systems); and f ðxÞ = mode shape normalized with respect to plitude-dependent behavior of the damping as plotted in Fig. 23.
maximum midspan deflection. The damping values for both Stages 1 and 2 are very similar and
Eq. (1) can also be written as show the same trend for the nonlinear damping phenomenon. It
is observed that for very low excitations, the damping ratio value
Meff ¼ f T M f (3) can go as low as 0.0045, and for increased excitations, the damp-
ing ratio value is around 0.0098.
where Meff = effective mass; M = lumped mass matrix caused by
the tributary area of each accelerometer point on the footbridge; Verification of Amplitude-Dependent Damping with
and, f = mode shape vector normalized with respect to maximum Acceleration Predictions in the FE Model
midspan deflection.
Accelerometer locations and the first bending mode shape were For vibration researchers, the natural frequency test data have been
shown in Fig. 21. The corresponding mass matrix is relatively easier to match with FE models than the acceleration

© ASCE 04016005-11 J. Archit. Eng.

J. Archit. Eng., 04016005


Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 22. MEScope mode shape animation

Table 4. Effective Mass Results via Mode Shapes for the First Bending Table 6. Stage 2: Calculated Damping Ratios via Effective Mass
Mode Procedure

Stage Effective mass calculations via MEScope mode shapes Sine excitation
Calculated damping
amplitude
1 0.470 M Acceleration response ratio ( b ) using
2 0.502 M N lbs amplitude (test) (%g) Meff = 0.502 M
33.8 7.6 0.12 0.00451
53.4 12.0 0.17 0.00503
Table 5. Stage 1: Calculated Damping Ratios via Effective Mass 120.1 27.0 0.24 0.00801
Procedure 195.7 44.0 0.32 0.00979
Sine excitation
Calculated damping
amplitude
Acceleration response ratio ( b ) usring
N lbs amplitude (test) (%g) Meff = 0.470 M FE models have always proved to be more difficult (Barrett 2006;
Davis 2008). Because the acceleration response predictions require
56.5 12.7 0.14 0.00690 the user to input the damping value for the FE model, the results are
118.8 26.7 0.26 0.00781 highly dependent on the often nonlinear damping value (Van
198.8 44.7 0.36 0.00944 Nimmen et al. 2014). Researchers usually use curve-fitted damping
ratios for acceleration prediction (Rainer 1979) and FE modeling
purposes (Živanovic et al. 2005, 2006).
response test data. Because the natural frequency and the mode The level of nonlinearity in damping has been shown in this pa-
shapes of a structure are independent of the loading and damping of per thus far. Yet, when the damping ratios for different levels of the
the structure, natural frequency predictions are easier to match with excitations are known, the FE models can be checked for the accel-
the FE models. In vibration serviceability research, the success rate eration responses.
of matching the acceleration response is not as high as matching the For that purpose, the sinusoidal excitations of Stages 1 and 2 are
natural frequencies with FE models. Acceleration predictions with run in the SAP2000 software with the calculated modal damping

© ASCE 04016005-12 J. Archit. Eng.

J. Archit. Eng., 04016005


ratios shown in Tables 5 and 6. The acceleration predictions by the success in matching the acceleration levels with various levels
SAP2000 are shown in Tables 7 and 8. It is clear in Tables 7 and 8 of excitations (and corresponding damping values) is unique to this
that the modal damping ratios introduced for different levels of paper. Successful verification and clarification of the amplitude-
excitations result in excellent correlation with the FE model acceler- dependent phenomenon and FE model matching of acceleration
ation predictions and collected accelerations data. The acceleration response develops confidence in the FE models not only for natu-
predictions are within 3.5% of the test data for Stage 1 and within ral frequency predictions but also for the acceleration predictions
2.5% for Stage 2. This is indeed a very successful set of acceleration in vibration serviceability applications.
predictions considering the fact that the FE models of the footbridge To develop practical applicability for the computed amplitude-
had already proved to be successful for static testing (deflections dependent damping values of this study, additional studies need
and load cell readings; Avci and Murray 2012) and natural frequen- to be conducted exploring a larger parameter space of foot-
cies (Avci 2014). Fig. 24 shows the summary of the tasks conducted bridges. With a larger number of experimental data, some confi-
for the tests and FE model simulations as explained in this paper. In dence can be developed in the amplitude-dependent damping
Fig. 24, the SAP2000 load case window for sinusoidal excitation is behavior in floor vibration serviceability. For that purpose,
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

for the loading amplitude of 198.8 N (44.7 lbs). The resulting accel- while reporting damping values in floor vibration testing, the ex-
eration response for this loading is also sinusoidal with an ampli- citation levels of the tests also need to be reported. Such future
tude of 0.348g in Fig. 24. The test data for this case are also shown studies would result in a data set that would offer greater insight
in Table 7. This is an excellent match because the test value for the into amplitude-dependent damping phenomenon in floor vibra-
acceleration response was about 0.36 g. The ratio for the SAP2000 tion applications.
acceleration prediction and the test acceleration are 0.348/0.36 =
0.967, within 3.5%.
As mentioned previously, one of the most difficult tasks in vibra- Conclusions
tion serviceability research is matching the test acceleration
responses with the FE model predictions. On top of the previous This paper presented a closer look into the amplitude-dependent
verifications of the FE model with static and dynamics test results, damping behavior in vibration serviceability. The work described
in the paper focuses on the nonlinear nature of damping by studying
0.012 a laboratory footbridge experimentally and analytically. The labora-
tory footbridge structure was studied extensively with static and
Calculated Modal Damping Ratio

0.010 dynamic tests. Three-dimensional FE models were developed,


0.008
updated, and fine-tuned for the two bottom chord extension configu-
rations for both static and dynamic tests.
0.006 The amplitude-dependent damping behavior of the footbridge is
shown for different amplitudes of sinusoidal excitations. Damping
0.004 Stage 2
ratios under different levels of excitations are verified with the FE
Stage 1
0.002 model acceleration predictions, which successfully matched the
acceleration responses from the tests when the corresponding modal
0.000
damping values are used from the resonant condition effective mass
0 10 20 30 40 50
calculations.
Sinusoidal Excitaon Amplitude (lbs) Although one of the most difficult tasks in vibration serviceabil-
ity research is matching the measured acceleration responses with
Fig. 23. Sinusoidal excitation amplitudes versus calculated modal
damping ratio through effective mass calculations
the FE models, the success of this paper in matching the accelera-
tion responses for various levels of excitations (using nonlinear

Table 7. Acceleration Responses (Tests and FE Model Predictions): Stage 1

Sine excitation
amplitude
Acceleration response Calculated damping ratio ( b ) Acceleration response Acceleration response amplitude
N lbs amplitude (test) (%g) using Meff = 0.470 M amplitude SAP2000 (%g) ratio: SAP2000/test
56.5 12.7 0.14 0.00690 0.135 0.964
118.8 26.7 0.26 0.00781 0.252 0.969
198.8 44.7 0.36 0.00944 0.348 0.967

Table 8. Acceleration Responses (Tests and FE Model Predictions): Stage 2

Sine excitation
amplitude
Acceleration response Calculated damping ratio ( b ) Acceleration response Acceleration response amplitude
N lbs amplitude (test) (%g) using Meff = 0.502 M amplitude SAP2000 (%g) ratio: SAP2000/test
33.8 7.6 0.12 0.00451 0.1226 1.022
53.4 12.0 0.17 0.00503 0.1736 1.021
120.1 27.0 0.24 0.00801 0.2453 1.022
195.7 44.0 0.32 0.00979 0.3271 1.022

© ASCE 04016005-13 J. Archit. Eng.

J. Archit. Eng., 04016005


Test the footbridge with the sinusoidal
excitaon at the natural frequency.
ap
p0

System Response
Input Excitation
Input Output

Time Time

Modal Tesng Apply the same sinusoidal excitaon on the FE model


Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

Get the acceleraon


Obtain the
response in FE model
natural
frequency

FE Model Acceleraon Response =


134.6 in/sec2 = 0.348g

Sinusoidal Excitaon Amplitude = 44.7 lbs = 198.8 N

Input Modal Damping Rao = 0.00944

Fig. 24. Overall summary of the tasks

damping values) is unique. The previous success of the FE models References


for static testing and natural frequency predictions of the foot-
bridge structure has been elevated to a new level with successful Aquino, R., and Tamura, Y. (2013). “Framework for structural damping pre-
FE model matching of measured acceleration responses in addition dictor models based on stick-slip mechanism for use in wind-resistant
to verification and clarification of the amplitude-dependent phe- design of buildings.” J. Wind Eng. Ind. Aerodyn., 117, 25–37.
nomenon. This reinforces the already existing confidence in the Avci, O. (2014). “Modal parameter variations due to joist bottom chord
FE models in vibration serviceability research. It is shown that the extension installations on laboratory footbridges.” J. Perform. Constr.
Facil., 10.1061/(ASCE)CF.1943-5509.0000635, 04014140.
FE models are reliable not only for natural frequency predictions
Avci, O., and Murray, T. M. (2012). “Effect of bottom chord extensions on
but also for the acceleration predictions in structural vibration
the static flexural stiffness of open-web steel joists.” J. Perform. Constr.
serviceability. Facil., 10.1061/(ASCE)CF.1943-5509.0000262, 620–632.
Avitabile, P. (2008). “Modal space: My coherence is better in some meas-
urements than others when impact testing. Am I doing something
Acknowledgments wrong?” Exp. Tech., 32(4), 17–18.
Bachmann, H., et al. (1995). Vibration problems in structures, Birkhauser,
The author thanks NUCOR Research and Devel-opment for Basel, Switzerland.
funding the research reported here and the staff of Thomas M. Barrett, A. R. (2006). “Dynamic testing of in-situ composite floors and
Murray Structural Engineering Laboratory at Virginia Tech for evaluation of vibration serviceability using the finite element method.”
their assistance with construction of the footbridge and the Ph.D. dissertation, Virginia Polytechnic Institute and State Univ.,
physical testing. Blacksburg, VA.

© ASCE 04016005-14 J. Archit. Eng.

J. Archit. Eng., 04016005


Brownjohn, J. M. W., and Pavic, A. (2007). “Experimental methods for esti- Middleton, C. J., and Brownjohn, J. M. W. (2010). “Response of high fre-
mating modal mass in footbridges using human-induced dynamic exci- quency floors: A literature review.” Eng. Struct., 32(2), 337–352.
tation.” Eng. Struct., 29(11), 2833–2843. Morita, K. O. I. C. H. I., and Kanda, J. (1996). “Experimental evaluation of
Butterworth, J., Lee, J., and Davidson, B. (2004). “Experimental determina- amplitude dependent natural period and damping ratio of a multi-story
tion of modal damping from full scale testing.” Proc., 13th World Conf. structure.” Proc., 11th WCEE.
on Earthquake Engineering, Vancouver, Canada. Pavic, A. (1999). “Vibration serviceability of long-span cast in-situ concrete
Casini, P., Giannini, O., and Vestroni, F. (2013). “Effect of damping on the floors.” PhD thesis, Dept. of Civil and Structural Engineering, Univ. of
nonlinear modal characteristics of a piecewice-smooth system through Sheffield, Sheffield, U.K.
harmonic forced response.” Mech. Syst. Sig. Process., 36(2), 540–548. Racic, V., Pavic, A., and Brownjohn, J. M. W. (2009). “Experimental iden-
Chang, C. J., and Mohraz, B. (1990). “Modal analysis of nonlinear systems tification and analytical modelling of human walking forces: Literature
with classical and non-classical damping.” Comput. Struct., 36(6), review.” J. Sound Vib., 326(1), 1–49.
1067–1080. Rainer, J. H. (1979). “Dynamic testing of civil engineering structures.”
CSI (Computers and Structures, Inc.). (2002). “SAP2000 linear and nonlin- Proc., Third Canadian Conf. on Earthquake Engineering, 551–574.
ear static and dynamic analysis and design of three-dimensional struc- Rainer, J. H., and Van Selst, A. (1976). “Dynamic properties of Lions’ gate
tures- getting started.” Berkeley, CA. suspension bridge.” ASCE/EMD Specialty Conf. on Dynamic Response
Downloaded from ascelibrary.org by Qatar University Library on 02/26/16. Copyright ASCE. For personal use only; all rights reserved.

Daoulatli, M. (2010). “Rate of decay of solutions of the wave equation with of Structures, 243–252.
arbitrary localized nonlinear damping.” Nonlinear Anal. Theory Salyards, K., and Noss, N. (2014). “Experimental evaluation of the influ-
Methods Appl., 73(4), 987–1003. ence of human-structure interaction for vibration serviceability.” J.
Davis, D. B. (2008). “Finite element modeling for prediction of low fre- Perform. Constr. Facil., 10.1061/(ASCE)CF.1943-5509.0000436,
quency floor vibrations due to walking.” Ph.D. dissertation, Virginia 458–465.
Polytechnic Institute and State Univ., Blacksburg, VA. Sharma, A., Patidar, V., Purohit, G., and Sud, K. K. (2012). “Effects on the
De Silva, C. W., ed. (2005). Vibration and shock handbook, CRC Press, bifurcation and chaos in forced Duffing oscillator due to nonlinear
Boca Raton, FL. damping.” Commun. Nonlinear Sci. Numer. Simul., 17(6), 2254–2269.
Díaz, I. M., and Reynolds, P. (2010a). “Acceleration feedback control of Smith, R. (2001). “Changing the effective mass to control resonance prob-
human-induced floor vibrations.” Eng. Struct., 32(1), 163–173. lems.” Sound Vib., 35(5), 14–17.
Díaz, I. M., and Reynolds, P. (2010b). “On-off nonlinear active control of Spence, S. M. J., Bernardini, E., Guo, Y., Kareem, A., and Gioffrè, M.
floor vibrations.” Mech. Syst. Sig. Process., 24(6), 1711–1726. (2014). “Natural frequency coalescing and amplitude dependent damp-
Ewins, D. J. (2000). Modal testing: Theory, practice, and application, 2nd ing in the wind-excited response of tall buildings.” Probab. Eng. Mech.,
Ed., Research Studies Press Ltd., Baldock, Hertfordshire, U.K. 35, 108–117.
Gromysz, K. (2013). “Verification of the damping model vibrations of rein- Steinwolf, A., Schwarzendahl, S. M., and Wallaschek, J. (2014).
forced concrete composite slabs.” Procedia Eng., 57, 372–381. “Implementation of low-kurtosis pseudo-random excitations to compen-
Ho, C., Lang, Z.-Q., and Billings, S. A. (2014). “A frequency domain analy- sate for the effects of nonlinearity on damping estimation by the half-
sis of the effects of nonlinear damping on the Duffing equation.” Mech. power method.” J. Sound Vib., 333(3), 1011–1023.
Syst. Sig. Process., 45(1), 49–67. Sun, C. T., and Lu, Y. P. (1995). Vibration damping of structural elements,
Hudson, M. J., and Reynolds, P. (2012). “Implementation considerations for Prentice Hall, Upper Saddle River, NJ.
active vibration control in the design of floor structures.” Eng. Struct., Taillon, J., Legeron, F., and Prud'homme, S. (2012). “Variation of damping
44, 334–358. and stiffness of lattice towers with load level.” J. Constr, Steel Res., 71,
Ingólfsson, E. T., Georgakis, C. T., Ricciardelli, F., and Jönsson, J. (2011). 111–118.
“Experimental identification of pedestrian-induced lateral forces on Tamura, Y., and Suganuma, S. (1996). “Evaluation of amplitude-dependent
footbridges.” J. Sound Vib., 330(6), 1265–1284. damping and natural frequency of buildings during strong winds.” J.
Inman, D. J. (2000). Engineering vibration, 2nd Ed., Prentice Hall, Upper Wind Eng. Ind. Aerodyn., 59(2–3), 115–130.
Saddle River, NJ. Tilly, G. P., Cullington, D. W., and Eyre, R. (1984). “Dynamic behaviour of
Jang, T. S. (2011). “Non-parametric simultaneous identification of both the footbridges.” IABSE Surveys S-26/84, No. 2/84, International Association
nonlinear damping and restoring characteristics of nonlinear systems of Bridge and Structural Engineering, Zurich, Switzerland, 13–24.
whose dampings depend on velocity alone.” Mech. Syst. Sig. Process., Van Nimmen, K., Lombaert, G., De Roeck, G., and Van den Broeck, P.
25(4), 1159–1173. (2014). “Vibration serviceability of footbridges: Evaluation of the cur-
Jang, T. S. (2013). “A method for simultaneous identification of the full non- rent codes of practice.” Eng. Struct., 59, 448–461.
linear damping and the phase shift and amplitude of the external harmonic Vibrant Technology, Inc. (2003). ME’ScopeVES online help, Scotts
excitation in a forced nonlinear oscillator.” Comput. Struct., 120, 77–85. Valley, CA.
Jeary, A. P. (1986). “Damping in tall buildings—A mechanism and a pre- Wu, J. R., Liu, P. F., and Li, Q. S. (2007). “Effects of amplitude-dependent
dictor.” Earthquake Eng. Struct. Dyn., 14(5), 733–750. damping and time constant on wind-induced responses of super tall
Jeary, A. P. (1996). “The description and measurement of nonlinear damp- building.” Comput. Struct., 85(15–16), 1165–1176.
ing in structures.” J. Wind Eng. Ind. Aerodyn., 59(2–3), 103–114. Wyatt, T. A. (1977). “Mechanisms of damping.” Symp. on Dynamic
Jeary, A. P. (1997). “Damping in structures.” J. Wind Eng. Ind. Aerodyn., Behaviour of Bridges, Transport and Road Research Laboratory,
72, 345–355. Crowthorne, U.K., 10–21.
Jeary, A. P., and Ellis, B. R. (1981). “Vibration tests of structures at varied Zapico-Valle, J. L., García-Dieguez, M., and Alonso-Camblor, R. (2013).
amplitudes.” Proc., 2nd Specialty Conf. on Dynamic Response of “Nonlinear modal identification of a steel frame.” Eng. Struct., 56,
Structures: Experimentation, Observation, Prediction and Control, 246–259.
ASCE, Reston, VA, 281–294. Zhang, W., and Zhang, H. (1994). “Modeling and analysis of nonlinear
Jeary, A. P., and Ellis, B. R. (1983). “On predicting the response of tall build- damping mechanisms in vibrating systems.” Int. J. Mech. Sci., 36(9),
ings to wind excitations.” J. Wind Eng. Ind. Aerodyn., 13(1–3), 173–182. 829–848.
Leonard, D. R., and Eyre, R. (1975). “Damping and frequency measure- Živanovic, S., Pavic, A., and Reynolds, P. (2005). “Vibration serviceability
ments on eight box girders.” Rep. No. LR682, Transport and Road of footbridges under human-induced excitation: A literature review.” J.
Research Laboratory, Dept. of the Environment, Crowthorne, U.K. Sound Vib., 279(1–2), 1–74.
Li, Q., Liu, D., Fang, J., Jeary, A., and Wong, C. (2000). “Damping in build- Živanovic, S., Pavic, A., and Reynolds, P. (2006). “Modal testing and FE
ings: its neural network model and AR model.” Eng. Struct., 22(9), model tuning of a lively footbridge structure.” Eng. Struct., 28(6),
1216–1223. 857–868.
Li, Q. S., Yang, K., Wong, C. K., and Jeary, A. P. (2003). “The effect of am- Zoghaib, L., and Mattei, P.-O. (2014). “Time and frequency response of
plitude-dependent damping on wind-induced vibrations of a super tall structures with frequency dependent, non-proportional linear damping.”
building.” J. Wind Eng. Ind. Aerodyn., 91(9), 1175–1198. J. Sound Vib., 333(3), 887–900.

© ASCE 04016005-15 J. Archit. Eng.

View publication stats J. Archit. Eng., 04016005

You might also like