You are on page 1of 12

2010 Structures Congress © 2010 ASCE 554

Finite Element Modeling of Common Lateral Systems in Tall Buildings: Insights


from Full-Scale Monitoring

Audrey Bentz1, Bradley Young2, Tracy Kijewski-Correa3, and Ahmad Abdelrazaq4


Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

1
Graduate Student, University of Notre Dame, Notre Dame, IN, USA
2
Structural Engineer, Skidmore, Owings, and Merrill, Chicago, IL, USA
3
Associate Professor, University of Notre Dame, Notre Dame, IN, USA
4
Samsung C&T/High Rise & Complex Buildings, Seoul, Korea

ABSTRACT

Tall buildings are one of the few constructed facilities whose design relies solely
upon analytical and scaled models, which, though based upon fundamental mechanics
and years of research and experience, lack thorough full-scale validation. In response
to this deficiency, the Chicago Full-Scale Monitoring Program was instituted and has
been continuously monitoring the responses of three tall buildings in Chicago since
2002, expanding recently to additional buildings internationally. Observations made
through this program, while showing general agreement between measured and
predicted behaviors, have also highlighted some discrepancies between the projected
periods of vibration using finite element models and the in-situ periods drawn from
full-scale data. The implications with respect to predicted accelerations are presented
in this study, various techniques for modeling elements of the lateral systems of these
buildings are explored, and sensitivity analyses are conducted to direct the physically-
meaningful calibrations of these finite element models to the full-scale observations.

INTRODUCTION

Even though high-rise construction serves as one of the most challenging projects
undertaken by society each year, tall buildings are one of the few constructed
facilities whose design relies upon analytical and scaled models, which, though based
upon fundamental mechanics and years of research and experience, have received
limited systematic full-scale validation. Since the performance of constructed
buildings directly affects public safety and comfort, full-scale validation of design
practice is of great interest to structural engineers as an increasingly practical and
viable means to assess performance. While strength design is the most important
aspect of any project, habitability criteria often govern the design of tall, slender
structures, as excessive acceleration levels have been shown to negatively impact
occupant comfort (Bentz and Kijewski-Correa, 2009). In particular, the ability to
reliably assess habitability relies upon the accurate prediction of a tall building’s
dynamic properties and this ability is best verified through full-scale monitoring

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 555

programs, such as the Chicago Full-Scale Monitoring Program (CFSMP) established


in 2001 (Kijewski-Correa et al., 2006).

Because damping in particular is very difficult to predict in design and significantly


influences building accelerations, much of the efforts of this monitoring program
have focused on this elusive property. However, full-scale observations have also
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

underscored discrepancies between the periods of vibration projected using finite


element (FE) models and the in-situ periods drawn from full-scale data. While the
prediction of natural periods is generally straightforward, common engineering
assumptions may not be appropriate for some of the more complex structural
systems, and the errors in predicted periods that result from these assumptions can
affect the accuracy of survivability and habitability responses.

Previous studies have underscored modeling issues specific to certain structural


systems. For example, steel moment resisting frames are commonly modeled by
centerline to centerline “stick” elements, which result in an unconservative estimate
of overall frame flexibility due to significant deformations in the panel zone region of
the actual building. On the other hand, studies of reinforced concrete buildings have
shown that simplified modeling of girders, columns, and shear walls significantly
underestimates structural flexibility (Kim et al., 2009). Similarly, a study conducted
on three reinforced concrete buildings in Korea revealed the need to vary the stiffness
of floor slabs contributing to lateral resistance, depending on the degree of shear wall
participation (Yoon et al., 2004; Erwin et al., 2007). The current study looks at some
of these common modeling assumptions and their implications on FE model
calibration for the buildings of the CFSMP, which employ steel tube and concrete
core and outrigger systems.

OVERVIEW OF THE BUILDINGS

The buildings of the CFSMP, referenced herein as Buildings 1, 2, and 3, are


representative of structural systems common to high rise design. These buildings are
located in the same general locale of downtown Chicago. In terms of dynamic
characteristics, all three buildings are relatively stiff in torsion, though Buildings 2
and 3 do exhibit coupling between its modes. A brief description of noteworthy
features of each building’s structural system is now provided, noting that the owners
have requested anonymity for their buildings so some specifics are not disclosed:

Building 1: The primary lateral-load resisting system of this office and residential
building features a steel tube comprised of exterior columns, spandrel ties and
additional stiffening elements to achieve a near uniform distribution of load on the
columns across the flange face, with very little shear lag. As such, lateral loads are
resisted primarily by cantilever action, with axial deformation and very little shear
deformation. The tower is also founded on rock caissons that are concentrated at each
of the tower main columns (Kijewski-Correa et al., 2006).

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 556

Building 2: In this reinforced concrete office building, shear walls located near the
core of the building provide the primary lateral load-resistance. At two levels, this
core is tied to the perimeter columns via reinforced concrete outrigger walls to
control the wind drift and reduce overturning moment in the core shear walls
(Kijewski-Correa et al., 2006). The core wall system resists most of the shear forces,
while the exterior columns deform axially due to engagement by the outrigger walls,
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

which experience combined bending and shear, a deformation mode similarly


experienced by the core’s link beams. The tower is founded on caissons.

Building 3: The steel moment-connected, framed tubular system of this office


building is comprised of closely spaced, wide columns and deep spandrel beams
along multiple frame lines. Deformations of the structure are due to a combination of
axial shortening, shearing and flexure in the frame members, and beam-column panel
zone distortions (Kijewski-Correa et al., 2006), though the tower’s shear deformation
contributes approximately 50% of the total deformation. The tower is founded on
rock caissons.

A DESIGNER’S PHILOSOPHY ON FEM


While there are some general modeling assumptions that can be considered ‘industry
standard’ assumptions among Structural Engineers, there is also considerable
variation in these modeling assumptions, both on the individual element level as well
as on a ‘building systems’ level, across design offices and in the manner in which
these models are analyzed based on the commercial software employed. In particular,
the rapid development of a large number of commercial FE software packages has
created these nuances in the features and analyses employed by each package, driven
largely by their target application. Additionally, although the FE method is typically
included in the curriculum of advanced engineering degree programs, FE modeling
specifically of tall building systems is generally not and is instead guided by the
experience of the Structural Engineers within a particular design office. This
variability in FE modeling assumptions and packages certainly does not constitute an
industry flaw – the combination of experience and independent calculations ensure
that the analytical model is performing appropriately and giving rational results.
However, the recent advancement of architectural modeling tools that make possible
the development of complex geometric forms has increased the Structural Engineer’s
dependency on the capabilities of the rapidly expanding FE software in order to
evaluate and rationalize these structural systems. This industry trend heightens the
need for Structural Engineers to understand the nuances of each FE package and the
applicability of basic modeling assumptions, as now discussed.

Model Complexity
Simple techniques are useful at the project onset. For example, gravity elements are
commonly omitted when evaluating lateral resistance, which generally leads to an
overestimation of the building period that is likely to be slightly conservative in
determining the actual building lateral stiffness and dynamic properties. However,

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 557

present computer capabilities can now enable very detailed and accurate analysis
models that include gravity elements, though even when gravity load resisting
elements are included, the FE model should always remain as simple as possible to
allow clear evaluation of the structural elements of interest, ensure the intended load
path, and be conducive to modification for parametric studies.
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

Floor Diaphragms
For the common case of multi-story buildings that rely upon the in-plane stiffness of
the structural floor slab to distribute the applied lateral loads to the lateral load
resisting system, it is common practice to create a more computationally efficient
model by simply constraining lateral movements of elements at each elevation by
omitting the floor slabs and employing rigid diaphragms. Such use of fully rigid
diaphragms are generally thought to be very reasonable in most cases; however, this
is not always the case, e.g., in tall building systems that utilize the floor slab to
distribute lateral loads between more than one lateral load resisting system. In
particular, floor slabs are expected to experience stiffness reductions due to flexural
cracking under gravity loads and temperature and shrinkage cracking due to the
restraining effects. In such cases, it may be prudent to perform sensitivity studies with
varying levels of assumed diaphragm stiffness in order to understand the potential
variation in the load distribution between the lateral load resisting elements and the
mechanism of load transfer to allow for better understanding of the detailing and the
connectivity required.

Further, caution should be taken in the use of fully rigid diaphragms in the case of
atypical floor slab geometries (i.e., slabs in an ‘L’ configuration or with
configurations involving re-entrant corners or large openings), where the artificial
horizontal constraint may significantly overestimate the actual in-plane stiffness of
the floor slab or mask regions of high stress associated with the unique geometry of
the floor slab. Caution should also be taken in the use of fully rigid diaphragms for
slabs associated with direct lateral load transfer structures in tall building systems
such as outrigger walls and belt walls, as the effectiveness of the load transfer system
may be overestimated as a result of the assumed fully rigid in-plane stiffness of the
diaphragm slab. Moreover, using diaphragm in braced frame systems is also not
appropriate at tie levels as it will eliminate the horizontal tie deformation and
underestimate the behavioral response under gravity loads.

Element Cracked Stiffness Properties


In order to adequately assess the behavior and load distribution of reinforced concrete
or composite building structures in FE models, the estimated post-cracking stiffness
of the structural elements at the limit state under consideration must be employed. It
is common practice to base the expected level of member cracking on the member
stresses associated with the design loads, even for acceleration evaluation, though this
is an uncertain process. Generally, in reinforced concrete buildings, once cracking
occurs, it will not be possible to recover the stiffness fully, thus a conservative level
of cracking is assumed due to both gravity and lateral loads.

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 558

Member Connectivity
FE models represent individual beam and column members via one-dimensional
‘stick’ elements connected in a centerline-to-centerline manner. The actual geometry
of the joints is not considered in the stiffness representation of the members, as each
‘stick’ element is free to deform along its full length, including that portion of the
element that may be within the physical joint between two or more members. Many
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

commercially available FE programs enable joint ‘offsets’ to represent the physical


geometry of the member joints and allow definition of an associated stiffness for the
joint offset region. In reinforced concrete building structures, it is common practice to
define such member offsets at beam-to-column joints with an associated joint
stiffness that is fully rigid (no shear or flexural deformations are thus allowed to take
place within this region). This is thought to generally account for the high degree of
material confinement, and thus increased stiffness, at monolithically cast reinforced
concrete joints. However, it has been shown that in the case of steel moment frame
structures, deformations occurring within the panel zone region of the columns
(defined by the region within the beam depth at beam-to-column connections) can be
significant and should therefore be considered in the analysis and design of such
structures. This is true even when members remain elastic (Krawinkler, 1978;
Charney and Downs, 2004). Panel zone deformation can be a significant source of
system deformation in frames with closely spaced columns and deep spandrel beams,
where the panel zone regions represent a significant portion of the frame geometry.
In such cases, even the basic centerline-to-centerline ‘stick’ representation of the
member connectivity, without the implementation of end offsets, can give overly
conservative estimates of the overall system stiffness.

The subsequent discussion expands upon some of these issues, as in-situ periods for
buildings associated with the CFSMP are compared with those predicted from FE
models. This process begins by first explaining the initial modeling approaches used
for the buildings in this study.

INITIAL MODELING APPROACH

For tall, slender buildings and those lightly damped, motion perception often
becomes the governing design criteria. Therefore, it is critical that the engineer be
able to accurately predict the full-scale behavior of the structure, a capability that
implicitly relies on the ability to construct an accurate analytical representation
through a finite element model. For the buildings associated with this study, finite
element models were developed using popular commercial software: ETABS and
SAP 2000. The mass associated with the self-weight of the structure and the full-
weight of the exterior cladding system are included in the dynamic analysis.
Additionally, special attention is paid to the use of the building and the resulting
loading conditions at each floor in order to determine what fraction of the design
imposed load to include in the mass calculations for the dynamic analysis. Due to the
heights of the study buildings, the analysis includes the effects of building
displacement on the frequencies through a second-order (P-Δ) iterative analysis. The
buildings were modeled as fully fixed at the base, as this is thought to approximate

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 559

the generally high soil-structure interfacial stiffness observed in Chicago for


buildings under wind-induced lateral loads (Kijewski-Correa et al., 2006). However,
in some cases, especially for such tall and slender towers, the foundation flexibility
may not be negligible.

For Buildings 1 and 3, framed primarily in structural steel, the representation of the
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

member stiffness was straightforward, as the steel elements remain elastic at service
level loadings. For the reinforced concrete building (Building 2), adjustments were
made to selected lateral-load resisting elements to represent the post-cracking
stiffness of these elements under service level loads. Specifically, the flexural and
shearing stiffnesses of the link (coupling) beams within the shear wall system were
reduced to one-half and one-fifth of the elastic stiffness, respectively. The beam-
supported slab was modeled using shell elements, whose flexural stiffness was set to
one-half of the elastic stiffness in order to approximate the post-cracking behavior of
the slab, which transfers flexure and shear between the perimeter columns and core
shear walls (Kijewski-Correa et al., 2006).

Figure 1 shows the mode shapes for each of the Chicago buildings, all normalized
with respect to the top floor displacement. Table 1 provides the predicted design and
full-scale periods for each of the building’s first three modes. Full-scale values
presented here are the average results of previous stationary analyses, all conducted
in the time domain for consistency (Kijewski-Correa et al., 2006; Kijewski-Correa
and Pirnia, 2007), and differences shown are with respect to the predicted design
values. As other publications have focused extensively on the discrepancies in
assumed and observed damping values, the focus of this study is to compare
predicted design and full-scale periods of vibration, noting the implications any
discrepancies may have on response predictions.

FIGURE 1 – NORMALIZED FUNDAMENTAL MODE SHAPES FOR BUILDINGS 1-3 (KIJEWSKI-CORREA


ET AL., 2006)

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 560

TABLE 1- FEM PREDICTED AND FULL-SCALE PERIODS OF VIBRATION

Building 1 Building 2 Building 3


Period (s) Mode 1 Mode 2 Mode 3 Mode 1 Mode 2 Mode 3 Mode 1 Mode 2 Mode 3
Predicted 7 4.9 2 6.7 6.4 4.6 7.7 7.6 4.5
Full-Scale 7.06 4.85 1.99 5.56 5.58 3.41 8.55 8.45 4.35
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

Difference 0.9% -1.0% -0.5% -17.0% -12.8% -25.9% 11.0% 11.2% -3.3%

IMPACTS ON RESPONSE PREDICTIONS

To demonstrate the impact that period estimates have on subsequent response


predictions from wind tunnel analyses, accelerations were predicted using the
methodology described in Bashor (2010), including a coupled response framework
introduced by Chen and Kareem (2005), for a mean hourly gradient winds of 14 m/s
from the north (Event 1) and 18 m/s from the west (Event 2). Acceleration responses
were predicted using both the design and in-situ periods for a fixed level of 1%
critical damping. Table 2 shows the comparison between the predicted accelerations
using the design periods and the in-situ periods for the alongwind and acrosswind
responses to each storm. It is interesting to note that the largest deviations in period
do not always result in the largest deviations in predicted acceleration responses.
These results demonstrate the impact frequency estimates can have on acceleration
predictions, perhaps most notably for Building 3 in the westerly event, where the
11% difference in frequency yields a 21% difference in accelerations. Thus the
general practice of underestimating the frequency, which from a stiffness perspective
may be viewed as conservative, may not be conservative with respect to accelerations
due to the frequency-dependence of the force spectra. Depending on the regime of the
spectrum occupied by the building’s fundamental mode, a stiffer in-situ frequency
may actually lead to an increase in the force spectrum’s magnitude and thereby the
response. Thus frequency (stiffness) influences multiple terms dictating the
acceleration response and this effect is not readily predicted without access to the
wind tunnel test data. This is demonstrated by the fact that the frequencies of both
modes of Building 3 were overestimated by about 11%, though this difference hardly
affects the response predictions in one mode, the other mode is affected by over 7%.

TABLE 2- ACCELERATION RESPONSES SIMULATED USING PREDICTED AND IN-SITU PERIODS

Across-Wind Response Along-Wind Response


RMS Acceleration RMS Acceleration
Design Design
Design In-Situ Response Design In-Situ Response
Building Frequency Frequency
(milli-g) (milli-g) Difference (milli-g) (milli-g) Difference
Difference Difference
1 0.116 0.113 2.85% -0.9% 0.090 0.094 -4.35% 1.0%
Event 1 2 0.122 0.110 9.73% 12.8% 0.104 0.101 2.43% 17.0%
3 0.095 0.095 -0.46% -11.2% 0.091 0.088 3.30% -11.0%
1 0.167 0.165 0.92% 1.0% 0.155 0.144 7.09% -0.9%
Event 2 2 0.168 0.136 19.2% 17.0% 0.229 0.197 13.9% 12.8%
3 0.197 0.183 7.11% -11.0% 0.164 0.198 -20.9% -11.2%

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 561

TABLE 3- SUMMARY OF BUILDING-SPECIFIC SENSITIVITY STUDIES


Structural System Area of Investigation
Building 1 Stiffened Steel Tube None
Building 2 Reinforced Concrete Core/Outrigger Slab Stiffness, Structural System Characteristics
Building 3 Framed Steel Tube Panel Zones, Foundation Stiffness
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

MODEL INVESTIGATION

As Building 1 shows excellent agreement between predicted and in-situ periods of


vibration in all three modes, the FE model does not require any explicit calibration
and thus will not be discussed further. However, sensitivity studies were conducted
on Buildings 2 and 3 to uncover root causes for period discrepancies, as summarized
in Table 3.

Building 3

Building 3 demonstrates periods longer than predicted in the sway modes by 11%,
while the torsional mode was predicted quite reasonably. Considering the multitude
of variables that can affect the frequency of a built structure that are either unknown
or difficult to quantify, the magnitude of the discrepancy between the predicted and
measured frequencies for Building 3 is not surprising. However, the fact that the
measured frequencies were lower than predicted is somewhat unexpected and
warrants further investigation. A previous study by Kijewski-Correa et al. (2005)
discussed four possible sources for these discrepancies: Panel Zone (PZ) Stiffness,
Service Condition Mass Variability, Foundation Stiffness, and Beam/Column Frame
Connectivity. That study showed that the Beam/Column Frame Connectivity does
not significantly affect the predicted frequency. The same study also looked into the
Service Condition Mass Variability and determined that the discrepancy between
predicted and measured values should not be more than 4.5%. As a result, this study
will now explore the role of Panel Zone Stiffness and Foundation Stiffness.

Because frame shear and flexural deformations play a large role in the lateral
deformations of Building 3, and the beam-column joints consist of closely spaced
columns and deep spandrel beams, deformations within the panel zone can have a
significant affect on the overall stiffness of the structure. The FE model, initially
developed as part of this investigation, did not take into account these deformations;
the joints were instead modeled as completely rigid. This modeling approach can
result in unconservative predictions for such buildings. A previous study investigated
the role of panel zone shear deformations (Kijewski-Correa et al., 2007) using the
Scissors model (Charney and Marshall, 2006), which quantifies the elastic shear
stiffness of the panel zone and allows deformations to initiate outside of the panel
zone. To simulate this in the FE model, fully-rigid links were applied to all beam-
column joints that reflect the actual panel zone geometry. A rotational spring was
introduced at each joint, calibrated to reflect the panel zone shear stiffness, as well as
localized column flange bending at the point of connection with the beam flanges.
As the panel zone shear stiffness and localized column flange bending stiffness are

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 562

entirely a function of the member dimensional and material properties, a unique


rotational spring must be defined at all joints where these properties vary.

While the inclusion of panel zone shear deformations improved the predictions by 6%
and 3% in the first and second lateral modes, the overall stiffness is still
overestimated. Therefore, the panel zone’s flexural deformations are now considered.
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

Using equations for flexural panel zone deformations and modifying the moment of
inertia values (Downs, 2002), spring constants applicable to the Scissors model were
determined by following the method described in Charney and Downs (2004) for the
panel zone’s shear deformations. In order to reduce the rigidity of the joint from fully
rigid to allow for this added flexibility, these flexural PZ spring constants were added
in series to the panel zone shear spring constants determined by the previous study.
This method was verified for a beam-column subassembly by comparing the FE
model deflection under a lateral load to the theoretical deflections due to
beam/column flexural, shear, and axial behavior as well as the panel zone shear and
flexural behavior. The results of the panel zone shear model (PZMa) and the panel
zone shear and flexural model (PZMb) on Building 3 are shown in Table 4, along
with the results of a center-line model (CLM), the original prediction of a model with
partial rigid offsets (PRM), and the periods observed in-situ. Note that although the
panel zone studies produced an 8% and 4% improvement over the original
predictions in the first two sway modes, there is clearly another contributing factor,
particularly since the inclusion of panel zones adversely affects the predicted
torsional period by about 10%.

TABLE 4- FEM PREDICTED AND IN-SITU PERIODS OF VIBRATION FOR BUILDING 3

Predicted
In-Situ (s)
CLM (s) PRM (s) PZMa (s) PZMb (s)
Mode 1 8.0 7.7 8.2 8.3 8.6
Mode 2 7.9 7.6 7.8 7.9 8.5
Mode 3 4.7 4.5 4.9 5.0 4.4

In order to further improve the analytical predictions of Building 3 to more closely


match the actual frequencies of vibration, the modeling of the soil-structure interface
is now examined. While it was initially assumed a rigid boundary at the base, the
foundation’s axial flexibility can be explicitly modeled by replacing vertical restraints
with springs with stiffnesses of 90 ton/mm and 60 ton/mm, based on the experience
of the last author. As shown in Table 5, the periods of vibration understandably
increase as more flexibility is introduced to the FE model at the foundation level,
approaching the in-situ values, though accuracy in the torsional mode is again
compromised. As a result, more explicit modeling of the boundary conditions will be
the focus of a future study, in an effort to resolve the issues associated with the
torsional mode as well as the lateral modes.

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 563

TABLE 5- FEM PREDICTED PERIODS OF VIBRATION FOR BUILDING 3 WITH FLEXIBLE BASE

Spring Stiffness: 90 ton/mm 60 ton/mm


In-Situ (s)
PRM (s) PZMb (s) PRM (s) PZMb (s)
Mode 1 8.19 8.71 8.37 8.91 8.6
Mode 2 8.07 8.31 8.25 8.50 8.5
Mode 3
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

4.67 4.99 4.67 5.00 4.4

Building 2

Buildings with concrete as a primary material are often modeled with reduced
stiffness to represent cracking that will occur during the course of the structure’s life,
so periods shorter than predicted are expected to be observed, particularly early in the
structure’s life cycle. This certainly is the case in Building 2, which also manifests
sway modes that are much closer to each other in-situ than predicted. As the
discrepancies in period are direction-specific in this building, it is worth reiterating
that Building 2 features two different lateral load resisting systems in its two
orthogonal directions. In mode 1 (x-direction), lateral resistance relies on shear walls
and outriggers that are intended to generate more cantilever action; in mode 2 (y-
direction), lateral resistance relies on shear walls and link beams, assumed to have
appreciably more shear deformations. As mentioned previously, the degree of floor
slab participation in the lateral resistance, as well as the assumed cracking in the slab
and the link beams, is somewhat subjective and are now investigated. One may
suspect that these factors will have a greater influence for the second mode due to the
system’s reliance on these elements in this direction.

The stiffness properties of floor slab and link beams were first adjusted to pre-
cracking conditions (i.e., doubled from the initial settings used in the FE model), with
results shown in Table 6. The building model shows that the reduction in slab
stiffness actually had very minor effects on the lateral stiffness, affecting the second
mode (y-direction) where the halving of slab stiffness resulted in a 2.5% increase in
the period. The greater influence of slab in the y-direction is likely due to the greater
frame action of the lateral system in that direction, as previously speculated. Also
demonstrated in Table 6, halving the link beam stiffness in the initial model had
substantially greater effects in the y-sway and torsional modes, as one may expect,
elongating the period by 9.8% and 8.5%. It is possible that the assumed levels of
cracking in these link beams have not materialized at this point in the structure’s life
cycle. By using uncracked properties for both the slabs and link beams, the second
mode (y-sway) period is reduced by about 12% and is within 0.3% of the observed
value, indicating that the conservative modeling of the slab and link beam properties
are likely the source of discrepancy for this mode, whose structural behavior is more
reliant on these elements. The torsional mode is also significantly affected by these
parameters by about 10%, but is still substantially higher than observed. The first
mode is least affected by these parameters, which is expected due to the fact that the
lateral system relies on shear walls and outriggers. The modeling of the shear walls
and outriggers in particular will contribute significantly to both the x-sway (Mode 1)
and will be explored in a subsequent parametric study.

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 564

TABLE 6- FEM PREDICTED AND FULL-SCALE PERIODS OF VIBRATION FOR BUILDING 2

Case Mode 1- X (s) Mode 2- Y (s) Mode 3- T (s)


Initial FEM settings 6.52 6.40 4.44
(cracked slab and link beams)
Uncracked slab stiffness 6.46 6.24 4.38
Uncracked link beam stiffness 6.52 5.77 4.06
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

Uncracked slab & link beam stiffnesses 6.45 5.64 4.01


In-Situ 5.61 5.66 3.41

CONCLUSIONS
This study focused on the standard practices used in finite element modeling of tall
buildings, with emphasis on common modeling assumptions that often are not valid
for structures of this type. While it is generally held that modern finite element
capabilities can reliably predict periods of vibration, the present work demonstrates
how full-scale monitoring can provide important insights into the potential
discrepancies in periods and their influence on the predicted accelerations. In
particular, while it was noted that cantilever-dominated steel structures could be
represented quite reliably, steel structures with appreciably greater shear
deformations manifested longer in-situ periods. This study explored explicit treatment
of panel zone deformations, as well as foundation axial flexibility, and found that
while these may be viable sources of added flexibility for the sway modes of such
buildings, they had an adverse effect on the torsional mode. While discrepancies in
reinforced concrete systems are generally expected, due to assumed levels of cracking
that may have yet to be realized in the life cycle of the building, this study
underscored that the assumed levels of cracking in slabs had only modest effects on
the in-situ periods of the study building, with far more substantial influences due to
the assumed cracking in the link beams. This effect was particularly “directional,” as
one may expect, influencing only the modes reliant on the link beams (y-sway and
torsion). Still, even with these refinements, predicted torsional and x-sway periods are
still longer by 17% and 15%, respectively. The unresolved contributing factors in
both of these buildings now warrants additional sensitivity studies to achieve a
physically-meaningful model calibration, including more explicit modeling of the
foundations.

ACKNOWLEDGEMENTS
The authors wish to acknowledge the financial support of the National Science
Foundation, Grant CMMI 06-01143, as part of the Chicago Full-Scale Monitoring
Program (PI: Dr. Ahsan Kareem of the University of Notre Dame). The first author’s
support by the Schmitt Presidential Fellowship from the Graduate School of the
University of Notre Dame is also recognized. The authors also wish to thank Notre
Dame graduate student Rachel Bashor executing the response predictions.

Structures Congress 2010


2010 Structures Congress © 2010 ASCE 565

REFERENCES
Bashor, R., (2010). “Dynamics of Wind Sensitive Structures,” PhD Dissertation, Department of Civil
Engineering and Geological Sciences, University of Notre Dame, Notre Dame, IN (in progress)

Bentz, A. and Kijewski-Correa, T. (2009). “Wind-Induced Vibrations of Tall Buildings: The Role of
Full-Scale Observations in Better Quantifying Habitability,” IMAC XXVII: A Conference and
Downloaded from ascelibrary.org by PRINCETON UNIVERSITY LIBRARY on 06/16/13. Copyright ASCE. For personal use only; all rights reserved.

Exposition on Structural Dynamics, Orlando, Florida, February 9-12

Charney, F.A., and Downs, W.M., (2004). “Modeling Procedures for Panel Zone Deformations in
Moment Resisting Frames”, Connections in Steel Structures V: Innovative Steel Connections,
Amsterdam, the Netherlands, June 3-5

Charney, F., and Marshall, J. (2006). “A Comparison of the Krawinkler and Scissors Models for
Including Beam-Column Joint Deformations in the Analysis of Moment Resisting Steel Frames,”
Engineering Journal, First Quarter, 31-48

Chen, X., and Kareem, A., (2005). “Coupled Dynamic Analysis and Equivalent Static Wind Loads on
Buildings with Three-Dimensional Modes,” Journal of Structural Engineering, 131(7) 1071-1082

Downs, W., (2002). “Modeling and Behavior of Beam/Column Joint Region of Steel Moment
Resisting Frames,” M.S. Thesis, Dept. of Civil and Enviro. Eng., Virginia Tech, Blacksburg, VA

Erwin, S., Kijewski-Correa, T, and Yoon, S.W. (2007). “Full-Scale Verification of Dynamic Properties
from Short Duration Records,” Proc. of Structures Congress, Long Beach, CA, ASCE

Kijewski-Correa, T., Young, B., Baker, W.F., Sinn, R., Abdelrazaq, A., Isyumov, N. and Kareem, A.
(2005). “Full-Scale Validation of Finite Element Models for Tall Buildings,” Proc. of CTBUH 7th
World Congress, New York, NY, Oct. 16-19

Kijewski-Correa, T., Kilpatrick, J., Kareem, A., Kwon, D.K., Bashor, R., Kochly, M., Young, B.S.,
Abdelrazaq, A., Galsworthy, J., Isyumov, N., Morrish, D., Sinn, R.C., and Baker, W.F., (2006)
“Validating Wind-Induced Responses of Tall Buildings: Synopsis of the Chicago Full-Scale
Monitoring Program,” Journal of Structural Engineering, 132(10) 1509-1523

Kijewski-Correa, T., Pirnia, J.D., Bashor, R., Kareem, A., Kilpatrick, J., Young, B., Galsworthy, J.,
Isyumov, N., Morrish, D. and Baker, W. (2007). “Full-Scale Performance Evaluation of Tall Buildings
Under Winds,” Proc. of 12th Intl. Conf. on Wind Eng., July 2-6, Cairns, Australia, vol. 1, pp. 351-358.

Kijewski-Correa, T., and Pirnia, J.D. (2007). “Dynamic Behavior of Tall Buildings under Wind:
Insights from Full-Scale Monitoring,” The Structural Design of Tall & Special Buildings, 16, 471-486

Kim, Y.J., Yu, E., Kim, D.K., Kim, S.D. (2009). “Calibration of Analytical Models to Assess Wind-
Induced Acceleration Responses of Tall Buildings in Serviceability Level,” Engineering Structures,
31, 2086-2096

Krawinkler, H., (1978), “Shear in Beam-Column Joints in Seismic Design of Frames”, Engineering
Journal, 15 (3), American Institute of Steel Construction, Chicago, IL

Pirnia, J.D., Kijewski-Correa, T., Abdelrazaq, A., Chung, J., and Kareem, A. (2007). “Full-Scale
Validation of Wind-Induced Response of Tall Buildings: Investigation of Amplitude-Dependent
Dynamic Properties,” Proceedings of Structures Congress, Long Beach, CA, ASCE

Yoon, S.W., and Ju, Y.K. (2004). “Dynamic Properties of Tall Buildings in Korea,” Proc. of CTBUH
2004, Seoul, Korea, October 11-13

Structures Congress 2010

You might also like