You are on page 1of 10

Journal of Constructional Steel Research 103 (2014) 13–22

Contents lists available at ScienceDirect

Journal of Constructional Steel Research

Column demands in steel plate shear walls with regular perforations


using performance-based design methods
Hassan Moghimi, Robert G. Driver ⁎
Department of Civil and Environmental Engineering, University of Alberta, Edmonton, AB, Canada, T6G 2W2

a r t i c l e i n f o a b s t r a c t

Article history: The design force demands on the columns of steel plate shear walls tend to be very large. Several methods have
Received 11 February 2014 been proposed to reduce the moments and axial forces on the columns, while still achieving good system perfor-
Accepted 25 July 2014 mance. One such method is to reduce the strength of the infill plate by perforating it using a regular pattern of
Available online 15 August 2014
circular holes. The diameter of the holes selected and their center-to-center spacing control the decrease in lateral
shear strength of the system. An appropriate performance-based design method is proposed and used to study
Keywords:
Seismic design
the suitability of the perforated system. The results of a variety of finite element models showed that although
Shear walls introducing the perforations into the infill plates reduces the lateral strength of the system, the net demand on
Steel plates the columns can in some cases actually increase. The perforated system behavior is also sensitive to the pattern
Analysis of holes selected; a small change in the arrangement of the holes can have a significant effect on the moment
Steel columns demand on the columns. Also, the flexibility of the beam-to-column connections influences the sensitivity of
Finite element method the column design forces to the pattern of holes selected and affects the column demands considerably.
Perforations © 2014 Elsevier Ltd. All rights reserved.
Performance-based design

1. Introduction Another is to reduce the strength of the infill plate by introducing a reg-
ular pattern of perforations in the form of circular holes. In the latter
Steel plate shear wall (SPSW) systems have demonstrated excellent method, the decrease in shear strength of the system depends on the di-
performance as a lateral force resisting system. Experimental and nu- ameter of the holes and their center-to-center spacing. It has generally
merical studies have shown that they possess high shear strength and been intuitively assumed that there is a direct correlation between the
outstanding ductility and redundancy. When the system is designed reduction of lateral strength of the wall associated with the infill plate
with steel grades commonly available in North America, the high inher- perforations and the alleviation of column force demands. However, in
ent shear strength of the system often results in the need for only very most research programs on SPSWs with perforated infill plates, the lat-
thin infill plates that may cause handling and welding considerations eral performance of the wall has been investigated but little attention is
to govern the thickness chosen. According to the capacity design con- given to the internal force demands on the columns. In this study, the
cept, the selection of thicker infill plates than needed imposes larger focus is specifically on the nature of the demands on the columns of
moment and axial force demands on the columns of SPSWs. multi-story SPSWs with regular perforation patterns in their infill
Several methods have been proposed to reduce the excessive de- plates.
mands on the columns due to the panel over-strength. These methods
can be classified into two main approaches. In the first, the share of 2. Background
the column demands arising from the effect of frame action is reduced
by using semi-rigid beam-to-column connections, using reduced beam Roberts and Sabouri-Ghomi [1] conducted a series of tests on small-
sections with rigid beam-to-column connections, and/or using a hori- scale steel shear panels, each with a centrally-placed circular opening.
zontal strut at the mid-height of each story. In the second approach, The infill plates, referred to as shear panels, all had a 300 mm depth
the share of the column demands arising from the effect of tension with either a 300 mm or a 450 mm width. Two different shear panel
field action in the infill plates is reduced by decreasing the strength of thicknesses, equal to 0.83 mm or 1.23 mm, were selected for each of
the infill plates themselves. The use of light-gage, cold-formed or low- the panel sizes, making four different panel types. The thinnest panels
yield-point (e.g., 250 MPa) steel for the infill plates is one method. were aluminum alloy and the remainder of the materials were steel.
Only the modulus of elasticity and 0.2% offset yield stress were reported
⁎ Corresponding author.
for each infill panel material. Four different diameters for the central cir-
E-mail addresses: moghimi@ualberta.ca (H. Moghimi), rdriver@ualberta.ca cular opening were used in each panel type, making 16 different speci-
(R.G. Driver). mens in total. The selected hole diameters were equal to 0 (no opening),

http://dx.doi.org/10.1016/j.jcsr.2014.07.020
0143-974X/© 2014 Elsevier Ltd. All rights reserved.

Downloaded from http://www.elearnica.ir


14 H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22

60, 105, or 150 mm. Each panel was installed in a stiff, pin-ended story shear imposes large capacity force demands on the columns,
boundary frame and the frames were loaded at two opposite corners which leads to a negative impact on the economic competitiveness of
in the direction of the panel diagonal. Each specimen was tested the system. This research investigates the appropriateness of the infill
under quasi-static cyclic loading by applying at least four complete plate perforation method for the purpose of decreasing the demand
cycles of loading with gradually increasing peak lateral displace- on the columns.
ments. All the specimens exhibited ductile behavior with stable hys- Performance-based design principles are used to compare different
teresis loops. Based on the test results, the shear strength and lateral systems. First, a standardized seismic hazard-independent method is
stiffness of the perforated panels were approximated by applying a proposed to evaluate the target displacement based on the yield dis-
reduction factor, which is a linear function of the ratio of the hole placement of the system and the ductility-related force modification
diameter to the panel depth, to the corresponding properties of a factor. A numerical model then is developed and validated against the
solid panel. available test results. Three different four-story SPSW systems are con-
A quasi-static cyclic test on a single-story SPSW specimen with a sidered. One has no perforations, while the other two have different
perforated infill plate was conducted by Vian [2]. The frame centerline perforation patterns in the infill plates. Each wall is analyzed for two dif-
dimensions of the specimen were 2000 mm in height and 4000 mm in ferent types of beam-to-column connection: simple and rigid. The tar-
width. Reduced beam sections were used with rigid beam-to-column get displacements corresponding to ductility-related force
connections. The infill plate was fabricated from low-yield-strength modification factors equal to 2 and 5 are evaluated for each wall based
steel, with measured yield and ultimate stresses of 165 MPa and on the proposed method. The axial force and bending moment demands
305 MPa, respectively, and was 2.6 mm thick. The yield strength of the and the design beam–column interaction capacity for the columns are
frame members was nominally 345 MPa. A total of 20 holes of calculated. By comparing these values, the efficiency of perforation pat-
200 mm diameter were used in the infill plate in a staggered pattern terns for internal force reduction is investigated.
of four horizontal rows of five holes each. The holes’ centers were ar-
ranged at a 45° angle on a rectangular coordinate system with grids 4. Target displacement evaluation
300 mm apart along both the horizontal and vertical directions. The
specimen was tested under cyclic loading. Beam plastic hinges occurred In order to compare SPSWs with and without infill plate perfora-
at the reduced sections, and the specimen exhibited ductile behavior tions, criteria are needed to make this comparison possible. Although
with stable hysteresis curves. capacity design provides a reliable and rational design method, it is
Purba and Bruneau [3] investigated the behavior of perforated strength-based. As such, it does not provide any information regarding
SPSWs with a series of numerical models subjected to monotonic load- the lateral deformation distribution over the height of the system
ing. Models of individual perforated strips were studied and their load when it is under the effect of the design earthquake. Moreover, it cannot
versus elongation response was compared to the behavior of complete be used when different performance objectives are to be satisfied for
single-story models with the same overall dimensions and number of different seismic hazard levels. On the other hand, the performance-
holes as the wall tested by Vian [2]. Perforation diameters in the infill based design method eliminates these shortcomings. In this method,
plate from 10 mm to 300 mm were considered, and the plate thickness the performance objectives define the status of the building following
for all models was 5.0 mm. The researchers found that the results from a design earthquake that represents a desired seismic hazard level.
an individual strip analysis based on the perforation pattern geometry Hence, this method provides a design procedure that allows for the de-
can accurately predict the behavior of the perforated panel provided sign of lateral force resisting systems for different performance objec-
that the hole diameter is less than 60% of the overall strip width. It tives under different seismic hazard levels.
was suggested that the shear strength of the perforated SPSW panel In performance-based design, the target displacement, δt, which
can be evaluated by applying an empirical factor of 0.7 to the equation is typically the maximum roof displacement under the design earth-
proposed by Roberts and Sabouri-Ghomi [1] for centrally-placed circu- quake, δr, must be evaluated. Based on the Coefficient Method (ASCE
lar openings. [5]), the target displacement can be approximated as follows, in
Driver and Moghimi [4] studied the feasibility of various SPSW ar- terms of the fundamental natural mode of vibration in the direction
rangements for use in lower seismic regions. The lateral performance under consideration:
of different two-story SPSW systems, all with simple beam-to-column
connections, was evaluated with comprehensive finite element models.
δt ¼ δr ¼ Δe ψ ¼ φ1;r Г 1 Sd ðT 1e ; ξ1 Þψ ð1Þ
As part of the system optimization, several infill plate details were stud-
ied, including the potential inclusion of a uniform pattern of circular
perforations. In order to evaluate the demand on the compression col- where Δe is the roof displacement for a linear elastic multi-degree-
umn of the configurations studied, the applied horizontal and vertical of-freedom system, considering only the effect of the fundamental
forces from the infill plate adjacent to the first-story compression col- mode, φ 1,r is the ordinate of the normalized fundamental mode
umn and the total shear and axial force demands at the compression- shape at the target (roof) node (which is equal to unity), Г1 is the
column base were evaluated. As expected, the lateral shear strength of modal mass participation factor for the fundamental mode,
the wall with perforated infill plates was reduced considerably com- Sd(T1e,ξ1) is the spectral displacement at the effective fundamental
pared to one with a continuous plate, and the forces applied directly period (T1e) and its corresponding critical damping ratio (ξ1), and ψ
to the columns by the infill plates were also reduced. However, the is a correction factor that takes into account the effects of nonlinear
total shear force demand at the column base, and in some cases the response on the maximum lateral deformation of the system. The
moment action, in the columns actually increased. As a result of this subscript 1 signifies fundamental mode properties of the multi-
apparently paradoxical outcome, the current study aims to investigate degree-of-freedom system.
the performance of such systems – specifically in terms of column The effective fundamental period is the fundamental period of the elas-
demands – in greater depth. tic system, modified to account for yielding based on the idealized push-
over curve. It can be evaluated from T1e = T1√(Ki/Ke) = 2π √(M/Ke),
3. Scope and objectives where T1 is the fundamental period, M is the total effective seismic mass
of the system, Ki is the elastic lateral stiffness, and Ke is the effec-
One major parameter affecting the cost of SPSW systems is the col- tive lateral stiffness, which is a secant stiffness calculated at a
umn cross-section required. The panel over-strength due to the selec- base shear force of 60% (ASCE [5]) of the effective yield strength
tion of infill plates that are thicker than required to resist the design of the system.
H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22 15

The correction factor ψ takes into account the effects of yielding (co- displacement) of a multi-degree-of-freedom system under monotoni-
efficient Cy) and pinched hysteresis cycles and cyclic strength and stiff- cally increasing lateral displacement. The system is designed for a base
ness degradation (coefficient Cp) of the system on the maximum lateral shear of VD, and the effective yield strength and corresponding roof dis-
displacement. ASCE [5] defines this correction factor as the product of placement are Vy and Δy, respectively.
two factors as follows: ASCE [5] defines the ductility-related force modification factor, Rd,
h  i h  i based on a strength ratio equal to:
2 2 2
ψ ¼ C y C p ¼ 1 þ ðRd – 1Þ= a T 1e  1 þ ðRd – 1Þ = 800  T 1e
Rd ¼ V 1 =V y ¼ α 1 Sa W=V y ð4Þ
ð2Þ

where Rd is the ductility-related force modification factor, as defined in where V1 is the maximum elastic base shear for the fundamental mode,
the National Building Code of Canada (NBCC) [6] and NEHRP [7], and is a α1 = Me1/M is the effective mass ratio of the fundamental mode, Me1 is
measure of the extent of stable nonlinearity in the system. This factor the modal effective mass, and W = Mg is the effective seismic weight.
represents its capability to dissipate energy through nonlinear behavior, However, defining Rd this way does not guarantee that the displacement
and can be defined as the ratio of elastic base shear demand to the yield ratio, the ratio of target to yield displacement, will be proportional to the
strength of the system. A site class factor, a, is defined to take into ac- effective lateral stiffness Ke (Fig. 1). In this study, Rd is defined instead as
count the effect of soil conditions at the site of the building. Based on the following strength ratio, so that the effective stiffness of the system
site soil properties, ASCE [8] and NBCC [6] have separated potential is maintained:
site conditions into six different classes. It is recommended that where
the soil properties are not known in enough detail to determine the Rd ¼ V e =V y ¼ Δe= Δy ¼ φ1;r Г 1 Sd ðT 1e ; ξ1 Þ=Δy ð5Þ
site class, Site Class D be assumed. ASCE [5] defines a = 60 for Site
Class D. The correction factors Cy and Cp may be selected as 1.0 for
Substituting Eq. (5) into (1), the following equations for the target
periods greater than 1.0 s and 0.7 s, respectively. Also, for systems
displacement, δt, and associated ductility ratio, μt, are obtained:
with no degradation of stiffness and strength, the coefficient Cp can be
selected as 1.0. h i
Normally, for a given design case the spectral displacement for the δt ¼ Rd Δy C y C p ð6Þ
effective fundamental mode of vibration is evaluated based on following
equation:
h i
2
Sd ðT 1e ; ξ1 Þ ¼ Sa ðT 1e ; ξ1 Þg ðT 1e =2πÞ ð3Þ μ t ¼ δt =Δy ¼ Rd C y C p ð7Þ

where Sa(T1e,ξ1) is the design spectral acceleration for the effective fun- where Cy and Cp are defined in Eq. (2).
damental period and critical damping ratio of the structure in the direc- Defining Rd based on Eq. (4) (ASCE [5]) leads to a conceptually differ-
tion considered, and g is the acceleration due to gravity. ent target displacement than the one derived from Eq. (5). For the pur-
The target displacement can be evaluated by substituting pose of comparison, substituting Eq. (3) into (1) and replacing the term
Eqs. (2) and (3) into (1). This method is similar to the method proposed Sa g with its equivalent from Eq. (4), the term T1e/2π with √(M/Ke), and
by ASCE [5] for the nonlinear static analysis procedure. However, the the term Vy/Ke with Δy (Fig. 1), the following equation is obtained for the
method is site-dependent; that is, it depends on the seismic hazard target ductility ratio:
and site class at each location and is not appropriate for general compar-
isons of different systems. In this study, a standardized seismic hazard- h i
μ t ¼ Rd ðГ 1 =α 1 Þ C y C p ð8Þ
independent target displacement is proposed as a method of comparing
different systems. Fig. 1 shows the pushover curve (base shear vs. roof
For typical SPSWs, the values of Г1 and α1 are fairly constant and can
reasonably be selected as 1.3 and 0.7, respectively, as shown later
(Table 1). This makes the coefficient Г1/α1 equal to about 1.8. As such,
Eq. (8) yields about 80% larger values for the target displacement than
Eq. (7). Hence, the definition of Rd based on the force ratio (Eq. (4)) re-
sults in a decrease in effective lateral stiffness from Ke (Rd based on the
deformation ratio from Eq. (5)) to Kd (Rd based on the force ratio from
Eq. (4)), as depicted in Fig. 1.
Eq. (7) defines the ductility ratio mainly as a function of Rd. In this
study, for any desired level of ductility-related force modification factor,
the target displacement is evaluated from Eq. (6) and then the perfor-
mance of the wall is assessed for the ductility ratio obtained from
Eq. (7).

Table 1
Wall properties.

Wall α1 Γ1 Vy Δy Rd = 2.0 Rd = 5.0

μt δt δt/h μt δt δt/h

– – kN mm – mm % – mm %

A1 0.72 1.32 2646 24.1 2.2 51.8 0.72 6.5 156 2.2
A2 0.72 1.32 2638 24.3 2.2 52.3 0.73 6.5 157 2.2
B1/C1 0.74 1.33 1908 22.8 2.2 49.1 0.69 6.6 149 2.1
B2/C2 0.74 1.33 1815 22.0 2.2 47.2 0.66 6.5 143 2.0
Fig. 1. Pushover curve of a general multi-degree-of-freedom system.
16 H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22

5. Numerical model of steel plate shear wall system

In order to study the performance of different SPSW systems under


pushover analyses, a system that was tested previously and proved to
have very good response under cyclic loading has been selected. The
system is a four-story SPSW (Driver et al. [9]) that was tested under ver-
tical load concurrent with increasing cyclic lateral displacement. The
columns were spaced at 3.05 m center-to-center and the story height
was 1.83 m for the top three stories and 1.93 m for the first story. The
columns were W310 × 118 sections, the intermediate beams were
W310 × 60 sections, and the top beam was a W530 × 82 section. The
frame members met Class 1 (CSA [10]) and Highly Ductile Member
(AISC [11]) compactness requirements, and they satisfy the stiffness re-
quirements for boundary members of SPSWs (ωh and ωL requirements
(CSA [10]). The mean measured infill plate thicknesses were 4.54 mm,
4.65 mm, 3.35 mm, and 3.40 mm, respectively, for stories 1 (base) to
4 (top). Moment connections were used at the beam-to-column joints
and all the materials were hot-rolled steel. In the early stages of the
experiment force control was used and, after achieving the yield later-
al displacement for the first story, δy, the experiment switched to the
displacement control method. Three full cycles were conducted at
each deformation stage up to a deformation of 3δy, and two cycles
were used at each deformation step afterward. The system achieved
its maximum lateral strength at 5δy and its lateral resistance then de-
clined gradually to about 85% of the maximum value at 9δy. The roof
lateral displacement at yield, maximum base shear, and maximum
lateral deformation occurred at 32.0 mm, 92.5 mm, and 158 mm,
respectively.
The commercial general-purpose finite element code ABAQUS is
selected to model this wall. In order to verify the model, first the
wall is analysed under the same conditions as those used in the ex-
periment, but subjected to monotonic loading. As-built dimensions
are used; however, the fish plates are not modeled so the infill plates
are connected directly to the surrounding boundary elements. All
infill plates of the wall are modeled using four-node shell elements
with reduced integration (S4R element), and frame elements
(beams and columns) are modeled using two-node linear beam ele-
ments (B31). Material and geometric nonlinearities are both consid-
ered. The measured material properties for the beams, columns, and
infill plates are used. The von Mises yield criterion, along with the
Fig. 2. Comparison of numerical and test results for panel 1: (a) monotonic loading;
isotropic hardening rule for monotonic loading and kinematic hard- (b) cyclic loading.
ening rule for dynamic loading, is used.
Story shear versus story deflection relationships for panel 1 (first
story) from the analyses are compared with the experimental results
in Fig. 2. Fig. 2(a) shows the response under monotonic loading, 6. Steel plate shear wall systems investigated
where it can be seen that the model provides a good approximation to
the test response envelope up to the ultimate strength. However, it In this study, three different SPSW systems have been considered.
over-predicts the strength slightly at the early stage of yielding, which Wall A1, shown in Fig. 3(a), is the SPSW tested by Driver at al. [9] and
is mainly because the monotonic analysis disregards the effect of soften- studied in the previous section numerically. Wall A2 is similar to Wall
ing of material due to the cyclic nature of the loading applied to the test A1, but with simple shear beam-to column connections. Fig. 3(b) shows
specimen. Also, the model does not account for the gradual degradation Wall B1, for which all the properties are the same as Wall A1 except
of strength exhibited by the test specimen after a story drift of 5δy that it contains a regular perforation pattern in all four panels, as permit-
(lateral displacement 42.5 mm) because localized material failure and ted by CSA standard S16 [10] and AISC 341 standard [11]. There are 23
low-cycle fatigue failure of the system due to cyclic response is not holes in panels 1, 2 and 3, and 18 holes in panel 4. Wall C1, shown in
captured in the model. Fig. 3(c), is the same as Wall B1, but with a different arrangement of
Fig. 2(b) compares experimental results with the cyclic loading holes; there are 22 holes in panels 1, 2 and 3, and 18 holes in panel 4.
analysis results for two excursions at 4δy and two at 5δy. The ulti- Walls B2 and C2 are the same as Walls B1 and C1, respectively, but
mate load is accurately predicted by the model at each displacement with simple shear beam-to-column connections.
level and even at each excursion. Also, the model predicts the loading In all perforated walls, the holes are of equal diameter (D = 230 mm)
and unloading stiffnesses and the deflection at which the tension and regularly spaced vertically and horizontally (at a distance of
field in the panel becomes effective quite well. However, it over- 280 mm) over the entire area of the infill plates. The regular grid of stag-
predicts the shear strength of the panel slightly around the zero- gered holes allows the development of a diagonal tension field at 45°. The
displacement region. In general, the model predicts the monotonic shortest center-to-center distance between perforations is Sdiag =
and cyclic responses of the system up to its ultimate strength very 396 mm, which makes the ratio of D/Sdiag = 0.58 b 0.6. The distance be-
well, and a similar model is used for the pushover analyses discussed tween the first hole and the surrounding frame members is between
in the next sections. D = 230 mm and D + 0.7 Sdiag = 507 mm. Hence, the two
H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22 17

Fig. 3. SPSW systems studied: (a) solid panels; (b, c) perforated panels with different hole patterns.

perforation patterns satisfy all the requirements of CSA S16 [10] and deformation-controlled, while for greater axial loads, it is considered
AISC 341 [11] for perforated infill plates. The factored design shear force-controlled. Steel columns subjected to axial tension, with or with-
resistance of the perforated wall is equal to that of a conventional out bending moment, are considered deformation-controlled (ASCE
SPSW times the reduction factor 1-0.7D/Sdiag, which for the patterns [5]). Beam-to-column connections can be fully-restrained or partially-
selected is equal to 0.6, except that the clear distance between col- restrained. In this research, SPSWs A1, B1, and C1 all use fully-
umns is used instead of center-to-center. restrained moment connections, while in SPSWs A2, B2, and C2 the
connections are partially restrained.
7. Performance criteria
7.2. Capacities for different components and actions
7.1. Classification of structural components and actions
When a nonlinear pushover procedure is used, the component ca-
The acceptability of a particular force or deformation in an element pacities are checked against component demands calculated at the tar-
depends on the component type and the classification of the action to get displacement. Since deformation- and force-controlled actions are
which it is subjected. Each component is classified as primary or second- inherently different, component strengths are determined differently
ary. A structural component that is designed to resist earthquake effects for each. The component capacity under deformation-controlled actions
is classified as a primary component, while one that is not is classified as is the permissible inelastic deformation capacity considering all
secondary (ASCE [5]). As such, all components of a SPSW system are coexisting forces and deformations. Deformation capacities are defined
primary components. in the ASCE [5] provisions for different steel structural systems and are
Each action is classified as deformation-controlled (ductile) or force- usually expressed in terms of the yield displacement. For this case the
controlled (non-ductile) based on the characteristics of the strength expected strength is used, which is the mean value of the component
versus deformation curve for a component subjected to that action. If resistance under that action at the deformation level anticipated. Ex-
the elastic part of the curve is followed by plastic response in the form pected component strengths can often be evaluated by code design
of a yield plateau with significant residual strength, the action induces equations using expected material properties and taking the resistance
ductile behavior in the member and it is classified as deformation- factor equal to 1.0. Any reliable strain hardening may be incorporated in
controlled. However, if the elastic part of the curve is followed by a the component strength versus deformation curve. Conversely, the
rapid loss of strength, the action generates non-ductile behavior in the component capacity under force-controlled actions is the lower-
member and is classified as force-controlled (ASCE [5]). bound strength, considering all coexisting forces and deformations,
If the axial force in a member is less than 10% of its axial strength, which can typically be taken as the mean value minus one standard de-
the component is considered a beam; otherwise, it is a column. Flex- viation of the component strength for the action (ASCE [5]). This can
ure of beams is considered to be a deformation-controlled action, often be estimated by code design equations using lower-bound mate-
whereas axial compression of columns is considered force-controlled. rial properties and a resistance factor equal to 1.0.
Flexural loading of columns with a coexisting axial load at the target dis- The expected component strength of a steel beam or other
placement of less than 0.5PCL – where PCL is the lower-bound axial com- deformation-controlled flexural member is the lowest expected value
pressive capacity as defined in the next section – is considered obtained from the limit states of yielding, lateral–torsional buckling,
18 H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22

local flange buckling, and shear yielding of web. The lower-bound axial shows how the wall performs compared with designing it based on
compressive strength of a steel column or other force-controlled com- building design provisions for new construction.
pression member is the lowest lower-bound value obtained for the
limit states of column buckling, local flange buckling, and local web 8. Pushover analysis results
buckling. The expected tensile capacity of a steel column, with or with-
out moment, or other deformation-controlled tension member is its Pushover curves are developed assuming a uniform lateral load dis-
cross-sectional strength based on the expected yield stress of the tribution pattern over the height of each wall. A deformation-control
material (ASCE [5]). For steel members subjected to the combined ac- loading scheme is used, and the effects of gravity loads are considered
tions of axial compression and bending, the lower-bound strength is on the columns of each wall. The pushover curves for all SPSWs investi-
expressed in the form of a strength interaction ratio and is equal to gated are shown in Fig. 4. The difference between the curves of Wall A1
the largest ratio for the limit states of cross-sectional strength, overall and Wall B1 or C1 indicates that the proposed reduction factor based on
in-plane member strength, and lateral–torsional buckling strength. S16 (CSA [10]) and AISC 341 (AISC [11]), 1–0.7*D/Sdiag ≈ 0.6, provides a
Among the components of SPSWs, the columns – and especially the conservative evaluation for shear resistance of the perforated walls. At
compression column – are arguably the most critical elements. The the target displacement corresponding to Rd = 5 (δr ≈ 150 mm), the
compressive force in the critical column is usually very large and ac- shear resistance ratios of Walls B1 and C1 to Wall A1 are 0.76 and
counts for more than 50% of the axial compressive capacity of the mem- 0.79, respectively, and the same ratios for Walls B2 and C2 to Wall A2
ber. In such cases, the combined action of axial compressive force and are 0.67 and 0.71, respectively. In the discussions that follow, it is as-
bending moment is considered force-controlled. In the following sec- sumed that the pushover loading is left-to-right so that the columns
tion, the bending moment and axial compressive forces of both columns can be distinguished as “left” and “right” for clarity instead of “tension”
are investigated to assess the nature of column demands in SPSWs with and “compression”, since the sign of the axial force can change within
regular perforation patterns. the same column.

7.3. Comparison of ASCE [5] with design codes for new buildings 8.1. Column internal forces demands

ASCE [5] defines two Basic Safety Earthquake (BSE) hazard levels, Previous research has shown that the demand on columns of SPSWs
namely BSE-1 and BSE-2. The 5%-damped spectral response accelera- is complex and notably large. Column shear, moment, and axial force
tion for the BSE-2 earthquake hazard level has a 2% probability of demands result from a combination of the frame action of the boundary
exceedance in 50 years. This hazard level is consistent with the Maxi- members and tension field action in the infill plates. The level of the de-
mum Considered Earthquake (MCE) in NEHRP [7]. The 5%-damped mand depends on the geometry of the system (such as bay width and
spectral response acceleration for BSE-1 is defined as the smaller of story height, number of stories, and thickness of the infill plates), type
the spectral response for the earthquake hazard level with a 10% prob- of beam-to-column connection (rigid, semi-rigid, or simple), and me-
ability of exceedance in 50 years and two-thirds of the spectral re- chanical properties of the material, especially the infill plate. The effects
sponse for BSE-2. Design codes for new buildings in the U.S. usually of infill plate perforation on the column demands are studied here in the
utilize two-thirds of the 5%-damped spectral response acceleration for context of the performance-based design method. For this purpose, by
the MCE (BSE-2) hazard level as their seismic design loads (NEHRP assuming a desired level of Rd, the target displacement is evaluated
[7]). (NBCC [6] defines the MCE as the seismic design event.) However, from Eq. (6). Appropriate values for Rd could be selected as 2, 3.5, and
when the criteria of ASCE [5] are used for new buildings, slightly higher 5, corresponding to three ductility levels, such as limited-ductility (or
than the seismic rehabilitation goal of life safety performance levels for a Ordinary), moderately ductile (or Intermediate), and ductile (or Spe-
BSE-1 hazard level are needed in order to achieve a system comparable cial) SPSWs, respectively. At the target displacement, the deformation
with a similar system designed based on design codes. demand on members under deformation-controlled actions must not
In this research, the responses of different SPSW systems are inves- exceed the expected deformation capacity, and the force demand on
tigated for two seismic demand levels, defined earlier by ductility- members under force-controlled actions must not exceed the lower-
related force modification factors equal to 2 and 5. The deformation- bound strength.
controlled actions are then labeled as Immediate Occupancy, Life Safety, Table 1 shows the dynamic properties (α1 and Γ1), lateral yield
or Collapse Prevention performance levels. In each case, comparing the strength and displacement, and target (roof) displacement ductility
performance level with the description in the previous paragraph ratio, displacement, and displacement ratio (h = overall height of

Fig. 4. Pushover curves of the SPSW systems.


H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22 19

wall) for both Rd = 2 and Rd = 5 for the walls under investigation. The right column. Moreover, these figures show that although the differ-
values in Table 1 for Walls B1 and B2 are the same as for Walls C1 and ences between Walls B and C are apparently negligible, and they have
C2, respectively, since they are identical except for the perforation pat- the same target displacement and shear resistance, the differences in
tern. The table shows that the effective mass ratio and modal mass par- the right column moment demands near the intermediate floors for
ticipation factor of the fundamental mode are similar for all the walls the walls with simple connections are up to 20%, suggesting an inherent
and approximately equal to 0.7 and 1.3, respectively. Also, they have uncertainty in the response of perforated SPSWs that warrants further
similar target displacement ductility ratios for each level of the ductility study.
factor, Rd. However, since the perforated walls have a smaller yield dis- The higher moment demand in the columns of SPSWs with perforat-
placement compared with the wall with solid infill plates, they have a ed infill plates, despite the lower forces applied directly from the infill
smaller target displacement for the same level of the ductility factor. It plates, is attributed chiefly to two effects. First, the perforations cause
can be seen from the table that the target displacements for Rd = 2 an increase in the participation of frame action in the response of
and 5 are approximately 50 mm and 150 mm, respectively, for all the SPSWs with moment connections. Second, in both walls with simple
walls. Based on the results of the physical test by Driver et al. [9], the and rigid connections, the perforations decrease the lateral stiffness of
roof lateral displacement at the yield point, maximum base shear, and the infill plate and the wall in total, and that changes the lateral
maximum displacement occurred at 32.0 mm, 92.5 mm, and 158 mm, deformed shape of both columns over the height of the walls, having a
respectively. As such, the 50 mm roof displacement is between the significant impact on the moment distributions in both columns. All in
Immediate Occupancy and Life Safety performance levels, and the roof all, the column design demands for perforated walls could potentially
displacement of 150 mm is close to the Collapse Prevention perfor- increase in some cases – where much of the increased demand is locat-
mance level. However, the test results show the performance of the ed in the connection regions – as compared to an equivalent wall with
wall under cyclic loading, which causes considerably more tearing and solid infill plates, which is in contrast to the main idea behind the devel-
low-cycle fatigue failure in the system than monotonic loading. opment of the system. Moreover, the actual demands on the columns
The axial force and moment demands on both columns of each wall appear to be quite sensitive to the pattern of the perforations.
for Rd = 2 and Rd = 5 are shown in Figs. 5 and 6, respectively. As can be
seen from Figs. 5(a), (b), 6(a), and (b), the axial force demands for the 8.2. Column design checks
perforated walls are generally lower than for the solid walls. However,
Figs. 5(c), (d), 6(c), and (d) show that the column moment demands In order to study the actual column demands in detail, design equa-
for the perforated walls for both connection types tend to be larger tions are used that take into account the interaction of axial compression
than for the solid walls, notably at the floor levels and the base of the and bending moment. The lower-bound strengths of the columns are

Fig. 5. Column internal force demands for Rd = 2: (a) and (c) axial force and bending moment in left (tension) column, (b) and (d) axial force and bending moment in right (compression)
column.
20 H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22

Fig. 6. Column internal force demands for Rd = 5: (a) and (c) axial force and bending moment in left (tension) column, (b) and (d) axial force and bending moment in right (compression)
column.

calculated based on two design standards. S16 [10] defines the beam– to tensile loads and, as such, they are considered as fully deformation-
column strength ratio as the largest ratio obtained from the limit states controlled elements so no strength ratio is given. The strength ratios
of cross-sectional strength, overall in-plane member strength, and later- for the upper two stories of this column are typically very low. For the
al–torsional buckling strength. For doubly-symmetric rolled compact first two stories of the right columns, some of the beam–column
members subjected to flexure and compression, with moments primar- strength ratios are greater than 1.0, and they can be particularly high
ily about their major axis, AISC 360 [12] defines the strength ratio as at the target displacement consistent with Rd = 5. The reason is attrib-
either a combined approach by checking an interaction equation or tak- uted to the fact that the design equations are a conservative approxima-
ing the larger ratio for two independent limit states of in-plane instability tion of true capacities, and also the material models consider strain
and out-of-plane buckling or lateral–torsional buckling. Both methods hardening, while all the interaction equations are checked using yield
were checked in this research. values. The results show that the perforations tend to increase the
Tables 2 and 4 show the compression-force–moment interaction ra- strength interaction values in the upper stories of both columns and in
tios for Rd = 2 and Rd = 5, respectively, based on both S16 and AISC 360. the lower stories the improvement is mostly small.
The strength ratios presented are the maximum for the above- The results of this research are consistent with the results reported
mentioned limit states, which in almost all cases is that of the cross- by Driver and Moghimi [4] on different wall configurations simulated
sectional strength. The first two stories of the left columns are subjected using shell elements to model the boundary frame members. They

Table 2
Column strength ratios for Rd = 2.

Story Wall A1 Wall A2 Wall B1 Wall B2 Wall C1 Wall C2

S16 AISC S16 AISC S16 AISC S16 AISC S16 AISC S16 AISC

Left Column 1 – – – – – – – – – – – –
2 – – – – – – – – – – – –
3 0.20 0.13 0.21 0.13 0.32 0.27 0.34 0.29 0.30 0.24 0.32 0.27
4 0.20 0.14 0.22 0.17 0.25 0.21 0.34 0.30 0.24 0.19 0.31 0.26
Right Column 1 1.14 1.17 1.12 1.14 1.11 1.13 1.01 0.97 1.11 1.14 1.01 0.97
2 0.84 0.85 1.00 0.89 0.85 0.87 0.76 0.69 0.81 0.83 0.83 0.74
3 0.62 0.64 0.58 0.60 0.67 0.69 0.62 0.64 0.64 0.66 0.61 0.62
4 0.36 0.33 0.49 0.44 0.39 0.37 0.56 0.52 0.37 0.37 0.53 0.50
H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22 21

Table 3
Column deformation ductility ratios for Rd = 2.

Story Wall A1 Wall A2 Wall B1 Wall B2 Wall C1 Wall C2

θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd

Left Column 1 – 0.87c – 0.81c – 0.60c – 0.39c – 0.63c – 0.45c


2 – 0.09b – 0.09b – 0.03b – 0.00b – 0.04b – 0.02b
3 2.12c – 2.17c – 2.20c – 2.52c – 2.20c – 2.46c –
4 2.00c – 1.97c – 1.49c – 1.52c – 1.54c – 1.57c –
Right Column 1 FCa – FCa – FCa – FCa – FCa – FCa –
2 FCa – FCa – FCa –– FCa – FCa – FCa –
3 3.53c – 3.71c – 3.27c – 3.67c – 3.26c – 3.59c –
4 2.32c – 2.31c – 1.70c – 1.76c – 1.74c – 1.82c –
a
FC: Force-controlled action.
b
Immediate Occupancy performance level.
c
Life Safety performance level.

showed that although the perforated infill plate applies a smaller level demands at the beam-to-column joints, the expected ductility capaci-
of horizontal force to the column, the column base shear is actually larg- ties specified for moment frames provide conservative performance
er than in a similar wall with solid infill plates. The reason for this is at- evaluation metrics for SPSW systems.
tributed to the augmented frame action and change in lateral Table 3 shows that at Rd = 2.0, the bottom two stories of the right
deformation of the columns due to the perforations in the infill plates. columns of the different walls are all subjected to force-controlled ac-
Tables 3 and 5 show the deformation ductility ratios and associated tions, while the upper two stories resist deformation-controlled actions
performance levels of each column at each story for Rd = 2 and Rd = 5, at Life Safety performance level. The left columns are subjected to
respectively. For the first two stories of the left columns, which are sub- deformation-controlled actions at the Life Safety performance level,
jected to tension, a deformation-controlled action, the axial deforma- with the exception of the second story, which is at the Immediate Occu-
tion ductility ratios are shown in the table. The first two stories of the pancy performance level. Table 5 shows that at Rd = 5.0, again the bot-
right columns, which are subjected to an axial compressive force at tom two stories of the right columns of the different walls are subjected
the target displacement that exceeds 50% of the lower-bound axial com- to force-controlled actions, but the third and fourth stories resist
pressive strength, are force-controlled for both axial load and flexure, deformation-controlled actions at the Collapse Prevention and Life Safe-
and are identified in the table by the symbol FC. The top two stories of ty performance levels, respectively. The left columns are subjected to
both the left and right columns are subjected to deformation- deformation-controlled actions, mainly at the Life Safety performance
controlled actions, and the rotational deformation ductility ratio of the level, although other regions of the columns meet either the Collapse
member is presented. The rotational demand, θ, is calculated as a Prevention or Immediate Occupancy performance criteria. The results
chord rotation by dividing the relative lateral deformation in each show in general that the perforations provide either limited or no im-
story by the story height. The yield rotations of the columns are also de- provements to the performance of the columns under deformation-
fined based on the chord rotation. As proposed in ASCE [5], the point of controlled actions. The performance levels for Rd = 2 are essentially
contraflexure is assumed at the mid-length of the elements, and the cal- the same, since the perforations have only a small effect on the deforma-
culated rotation is then reduced linearly for the effect of axial force. tion ductility values, although some improvements are noted in the first
The deformation-controlled actions define the performance level of and fourth stories of the left column and the fourth story of the right.
the system by checking the deformation demand against the expected While some of the column regions also show performance improve-
deformation capacity. However, force-controlled actions must always ments at Rd = 5.0 and others are essentially equivalent, the perforations
satisfy the strength criteria for any performance level and seismic haz- increase the rotational ductility demand substantially in the third story
ard level. As such, the performance level of each column at each story of the left column of the walls with simple joints, changing the perfor-
is presented in Tables 3 and 5, wherever the columns are subjected to mance category of the wall from Life Safety to Collapse Prevention.
deformation-controlled actions (either axial or rotational). ASCE [5]
specifies the expected deformation capacities at each performance
level (Immediate Occupancy, Life Safety, and Collapse Prevention) 9. Conclusion
only for stiffened SPSWs, where the infill plates are sufficiently stiffened
to prevent buckling. Since the behavior of such a system is inherently In this paper, a standardized method has been presented to evaluate
different from unstiffened SPSWs, in the current study the acceptance the target displacement of any lateral force resisting system in terms of
criteria for steel moment frames are used to evaluate the performance the yield displacement of the system and the ductility-related force
levels of each column. Since the infill plate reduces the deformation modification factor, Rd. Also, a suitable method for comparing different

Table 4
Column strength ratios for Rd = 5.

Story Wall A1 Wall A2 Wall B1 Wall B2 Wall C1 Wall C2

S16 AISC S16 AISC S16 AISC S16 AISC S16 AISC S16 AISC

Left Column 1 – – – – – – – – – – – –
2 – – – – – – – – – – – –
3 0.25 0.19 0.35 0.31 0.67 0.72 0.49 0.47 0.64 0.66 0.47 0.45
4 0.23 0.18 0.35 0.31 0.44 0.43 0.53 0.52 0.42 0.40 0.51 0.50
Right Column 1 1.34 1.37 1.21 1.24 1.21 1.24 1.05 1.04 1.22 1.25 1.08 1.11
2 1.06 1.08 1.09 1.01 1.11 1.14 0.91 0.84 1.12 1.15 0.98 0.89
3 0.76 0.78 0.83 0.82 1.03 1.06 0.84 0.87 1.02 1.05 0.81 0.83
4 0.43 0.39 0.69 0.64 0.56 0.56 0.78 0.75 0.55 0.55 0.73 0.70
22 H. Moghimi, R.G. Driver / Journal of Constructional Steel Research 103 (2014) 13–22

Table 5
Column deformation ductility ratios for Rd = 5.

Story Wall A1 Wall A2 Wall B1 Wall B2 Wall C1 Wall C2

θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd θ/θy Δax/Δyd

Left Column 1 – 2.71c – 1.90c – 1.68c – 1.10c – 1.84c – 1.19c


2 – 0.03b – 0.06b – ≈0b – ≈0b – ≈0b – ≈0b
3 5.57c – 5.82c – 5.73c – 7.25d – 5.46c – 6.97d –
4 5.74c – 5.25c – 3.42c – 5.57c – 3.60c – 5.56c –
Right Column 1 FCa – FCa – FCa – FCa – FCa – FCa –
2 FCa – FCa – FCa – FCa – FCa – FCa –
3 9.96d – 10.81d – 9.59d – 10.97d – 9.29d – 10.70d –
4 6.71c – 6.34c – 4.04c – 6.59c – 4.24c – 6.57c –
a
FC: Force-controlled action.
b
Immediate Occupancy performance level.
c
Life Safety performance level.
d
Collapse Prevention performance level.

systems with the same value of Rd is provided, since the method is not Acknowledgements
seismic-zone dependent.
The concept of perforated SPSWs was introduced primarily to allevi- Funding for this research was provided by the Natural Sciences and
ate the severe force demands on the columns of SPSWs by degrading the Engineering Research Council. The test specimen was donated by Su-
lateral shear resistance of the system. This study shows that although preme Steel. Financial support for the first author was provided through
the perforations reduce the shear capacity of the infill plates consider- scholarships from the Steel Structures Education Foundation, Alberta
ably, and indeed do reduce the forces applied by the infill plates to the Heritage Foundation, University of Alberta, Canadian Society for Civil
columns, they may not reduce the net force and deformation demands Engineering, and Canadian Institute of Steel Construction. All support
on the columns. The results indicate that although the perforated infill is gratefully acknowledged.
plate consistently reduces the axial force demands in the columns, it in-
creases the contribution of frame action in the system and changes the
lateral deformation distribution of the columns. These effects increase References
the moment demands on the columns, especially in the upper stories.
[1] Roberts TM, Sabouri-Ghomi S. Hysteretic characteristics of unstiffened perforated
The increase in the moment is such that the perforations in some steel plate shear panels. Thin-Walled Struct 1992;14(2):139–51.
cases increase the design strength ratios for the columns based on the [2] Vian D. Steel plate shear walls for seismic design and retrofit of building structures.
combined force-controlled actions of axial compression and bending. [PhD dissertation] Buffalo, N.Y: State Univ. of New York at Buffalo; 2005.
[3] Purba R, Bruneau M. Finite-element investigation and design recommendations for
Also, in general the perforations provide little improvement to the per- perforated steel plate shear walls. J Struct Eng Am Soc Civil Eng 2009;135(11):
formance of elements under deformation-controlled actions. 1367–76.
The results of this research also show that a minor change in perfo- [4] Driver RG, Moghimi H. Modular construction of steel plate shear walls for low and
moderate seismic regions. Structures Congress, Structural Engineering Institute,
ration pattern may cause up to 20% differences in column moment
American Society of Civil Engineers, Las Vegas, NV; 2011. p. 758–69.
demand in some regions. This suggests an inherent uncertainty in the [5] ASCE. Seismic rehabilitation of existing buildings. ASCE/SEI 41–06. Reston, VA:
response of perforated SPSWs and the need to consider the specific lay- American Society of Civil Engineers; 2007.
[6] NRCC. National building code of Canada. Ottawa, ON, Canada: National Research
out of holes when determining column demands, rather than simply
Council of Canada; 2010.
using a uniformly distributed stress with a value reduced from that of [7] NEHRP. Recommended provisions for seismic regulations for new buildings and
a solid plate. other structures (FEMA 450). Part 2. Commentary. National Earthquake Hazards
It is emphasized that the research results presented in this paper do Reduction Program, Washington, DC2003 ed. ; 2004.
[8] ASCE. Minimum design loads for buildings and other structures. ASCE/SEI 7–10.
not imply that perforated infill plates cannot be used advantageously in Reston, VA: American Society of Civil Engineers; 2010.
the design of SPSWs. However, they have revealed behavioral complex- [9] Driver RG, Kulak GL, Kennedy DJL, Elwi AE. Cyclic test of a four-story steel plate shear
ities that require further study if the performance outcomes of these wall. J Struct Eng Am Soc Civil Eng 1998;124(2):112–20.
[10] CSA. Design of steel structures. CSA S16-14. Mississauga, ON: Canadian Standards
walls are to be evaluated with confidence. In particular, the effects of Association; 2014.
the perforations need to be investigated at different performance levels [11] AISC. Seismic provisions for structural steel buildings. ANSI/AISC 341–10. Chicago, IL:
in walls with different geometries, numbers of stories, panel aspect ra- American Institute of Steel Construction, Inc.; 2010.
[12] AISC. Specification for structural steel buildings. ANSI/AISC 360–10. Chicago, IL:
tios, perforation patterns, and boundary frame cross-sections. In addi- American Institute of Steel Construction, Inc.; 2010.
tion, seismic time history analyses are required to assess the dynamic
behavior of this type of SPSW.

You might also like