Cyclic Analysis of Wood Shear Walls

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

CYCLIC ANALYSIS OF WOOD SHEAR WALLS

By Bryan Folz1 and Andre Filiatrault,2 Member, ASCE

ABSTRACT: A simple numerical model to predict the load-displacement response and energy dissipation char-
acteristics of wood shear walls under general quasi-static cyclic loading is presented. In this model the shear
wall is composed of three structural components: rigid framing members, linear elastic sheathing panels, and
nonlinear sheathing-to-framing connectors. The hysteretic model for the sheathing-to-framing connector takes
account of pinching behavior and strength and stiffness degradation under cyclic loading. A robust displacement
control solution strategy is utilized to predict the wall response under general cyclic loading protocols. The shear
wall model has been incorporated into the computer program CASHEW (Cyclic Analysis of SHEar Walls). The
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

predictive capabilities of this program are compared with monotonic and cyclic tests of full-scale wood shear
walls. It is shown that this model can accurately predict the load-displacement response and energy dissipation
characteristic of wood shear walls under general cyclic loading. As an application of the CASHEW program, a
procedure is presented for calibrating a single degree-of-freedom system to predict the complete nonlinear
dynamic response of shear walls under seismic loading.

INTRODUCTION global wall response is fully attributable to the nonlinear load-


deformation behavior of the sheathing-to-framing connectors
In low-rise wood-frame structures subjected to earthquake (Tuomi and McCutcheon 1978; Gupta and Kuo 1985, 1987;
loading, shear walls are commonly used as the primary com- Filiatrault 1990). The sheathing is generally assumed to de-
ponent of the lateral load resisting system. Under such loading, velop only elastic in-plane shear forces. It was also found that
the racking response of a shear wall is generally cyclic in bending of the framing members contributed little to the global
nature. Building codes, however, have typically assigned de- wall response (Gupta and Kuo 1985). Consequently, in a num-
sign shear strength values to wood shear walls based on ex- ber of these studies the framing members are assumed to be
perimental data obtained from static racking tests. In recent rigid. These models generally provide good agreement with
years there has been a proliferation of full-scale testing per- the load-displacement response obtained from tests. However,
formed on wood shear walls under cyclic loading. The objec- because of their simplicity, they are not able to capture the
tive of this increased research activity has been to provide a detailed interaction and load sharing between the components
greater understanding of the response of shear walls under of the shear wall under the imposed lateral loading.
loading conditions that more closely resemble the seismic re- More sophisticated finite-element models have also been
sponse scenario. Cyclic loading, unlike monotonic loading, is proposed (Itani and Cheung 1984; Gutkowski and Castillo
not uniquely defined. Numerous cyclic loading protocols have 1988; Dolan and Foschi 1991; White and Dolan 1995). In
been proposed [European Committee for Standardization these models the framing members are composed of beam el-
(CEN) 1995; SEAOSC 1997; International Standards Orga- ements and the sheathing is represented by plane stress ele-
nization (ISO) 1999; Krawinkler et al. 2000], and the perfor- ments or plate bending elements. The sheathing-to-framing
mances of wood shear walls against these protocols have been connectors are modeled using springs with nonlinear load-de-
experimentally investigated (Rose 1994; Skaggs and Rose formation characteristics. Also, gap and bearing elements have
1996; He et al. 1998). Other experimental studies have focused been included along the interface between the sheathing pan-
on the degrading response of shear walls under cyclic loading els. Obviously, these models are able to capture more fully the
(Shenton et al. 1998). Still other experimental studies have intercomponent response within the wall. However, with this
considered the influence of panel size (Lam et al. 1997), open- increased model complexity, a greater computational effort
ings (He et al. 1999), fastener type and the contribution from must be expended. Interestingly, the overall global load-dis-
gypsum wallboard (Karacabeyli and Ceccotti 1996), and the placement predictions of these models produce essentially the
effect of hold downs (Commins and Gregg 1994) on the cyclic same level of correlation with experimental data as the simpler
response of wood shear walls. Even with this extensive effort models discussed previously.
in experimentally evaluating the cyclic response of wood shear Several nonlinear dynamic analysis models have also been
walls, there still remains a need for a more complete under- developed for predicting the seismic response of wood shear
standing of these structural elements. It is obvious that this walls. The simplest of these are single degree-of-freedom
cannot be achieved through testing programs alone; numerical (SDOF) lumped parameter models [e.g., Stewart (1987) and
studies should be conducted to complement this experimental Foliente (1995)]. These, however, have limited application be-
work. cause the model must be calibrated, in each case, to full-scale
A large number of numerical models of varying complexi- test data. Others have extended existing static shear wall mod-
ties have been formulated to predict the static racking response els to perform nonlinear dynamic time history analysis under
of wood shear walls. In the simpler models, the nonlinear seismic input (Dolan 1989; Filiatrault 1990).
1
Between the bookends of static ultimate load analysis and
Visiting Assoc. Prof., Dept. of Struct. Engrg., Univ. of California, San the full nonlinear dynamic analysis lies the cyclic analysis of
Diego, 9500 Gilman Dr., La Jolla, CA 92093-0085.
2
Prof., Dept. of Struct. Engrg., Univ. of California, San Diego, 9500 shear walls. As mentioned previously, it is this loading regime
Gilman Dr., La Jolla, CA 92093-0085 (corresponding author). E-mail: that has received the greatest experimental research attention
afiliatrault@ucsd.edu of late. Surprisingly, however, very little research has been
Note. Associate Editor: David Rosowsky. Discussion open until Sep- applied to developing structural analysis models that can com-
tember 1, 2001. To extend the closing date one month, a written request plement this experimental effort. The main objective of this
must be filed with the ASCE Manager of Journals. The manuscript for
paper is to present a formulation for the structural analysis of
this paper was submitted for review and possible publication on Septem-
ber 6, 2000; revised November 20, 2000. This paper is part of the Journal wood shear walls under general cyclic loading. The numerical
of Structural Engineering, Vol. 127, No. 4, April, 2001. 䉷ASCE, ISSN model presented herein predicts, for sheathed shear walls
0733-9445/01/0004-0433–0441/$8.00 ⫹ $.50 per page. Paper No. 22546. with or without opening, the load-displacement response and
JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001 / 433

J. Struct. Eng., 2001, 127(4): 433-441


energy dissipation characteristics under arbitrary quasi-static
cyclic loading. In formulating this structural analysis tool,
a balance has been sought between model complexity and
computational overhead. The proposed model is validated
against full-scale tests of wood shear walls subjected to mon-
otonic and cyclic loading. It is also shown how this model can
be used to calibrate the parameters of a SDOF system to pre-
dict the nonlinear dynamic response of wood shear walls to
seismic ground motions.

MODEL FORMULATION
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

Structural Configuration
Typical wood shear wall assemblies are composed of four
basic structural components, as shown in Fig. 1: framing mem-
bers, sheathing panels, sheathing-to-framing connectors, and
hold-down anchorage devices. The framing members are gen-
erally sawn lumber pieces oriented horizontally (plates and
sills) and vertically (studs) with only nominal nailing to hold
the framework together. Sheathing panels are usually made of
plywood, oriented strand board, or other structural panel prod-
ucts. These panels may be applied to one or both sides of the
wall. Openings in the paneling may occur to facilitate place-
ment of windows or doors. In such cases, additional framing
members are required to reinforce these openings. Sheathing-
to-framing connectors are most commonly dowel-type fasten-
ers such as nails. These fasteners are typically spaced at reg-
ular intervals, with fastener lines around the perimeter of the
sheathing panels more densely spaced than throughout the
panel interior. Hold-down devices, in the form of simple an-
chor bolts or proprietary hardware, may be included in the
wall assembly to facilitate shear transfer, limit global over-
turning under lateral loading, and provide additional connec-
tion between the sill and the perimeter stud framing members.

Kinematic Assumptions FIG. 2. General Deformation of Framing Members and Typical


Sheathing Panel in Shear Wall: (a) Undeformed Configuration;
Figs. 2(a and b) show the undeformed and deformed con- (b) Deformed Configuration (Racking Mode)
figuration (racking mode) of a typical shear wall under the
action of a prescribed lateral displacement UF applied at the
top of the wall. Under this action, the original orthogonal grid order effect on wall response (Gupta and Kuo 1985). Hence,
work of framing members distorts into a parallelogram with in this study the framing members are assumed to be rigid
the top plate and sill remaining essentially horizontal. It is with pin-ended connections, given the nominal attachment that
assumed herein that the sill is sufficiently anchored so that is prescribed for these members. As a consequence of these
uplift is effectively eliminated. Previous research has shown assumptions, the grid work of framing members alone is mod-
that the in-plane bending of framing members has a second- eled as a mechanism with no lateral stiffness. As seen from
Fig. 2(b), the lateral displacement of the framing can be char-
acterized by a SDOF; the top of frame lateral displacement or
drift UF.
As illustrated in Fig. 2(b), when the wall racks, each rec-
tangular sheathing panel develops a uniform in-plane shear
deformation Us superimposed on rigid-body translations Ū and
V̄ and rotation ⌰. These rigid body modes are each measured
with respect to the centroid of the panel (x̄, ȳ) in the unde-
formed configuration. It is assumed that the sheathing panels
have sufficient stiffness so that out-of-plane panel deforma-
tions can be ignored. Previous research supports this assump-
tion for typical sheathed shear walls (Dolan 1989). Thus, each
sheathing panel within the wall is assigned only four degrees
of freedom: Us, Ū, V̄, and ⌰.
At the ultimate lateral load-carrying capacity of a typical
wood shear wall, the corresponding lateral drift is of the order
of 2–4%. Consequently, in this study it is assumed that the
individual deformations of the framing members and sheathing
panels are relatively small. It then follows that a generic point
Q on the sheathing panel with initial local panel coordinates
(x, y), as shown in Fig. 2, experiences the following linearized
FIG. 1. Components of Typical Wood Shear Wall deformations up and vp under racking of the wall:
434 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001

J. Struct. Eng., 2001, 127(4): 433-441


¯ ⫹2
up = U 冉冊
y
h
Us ⫺ y⌰ (1)

¯ ⫹ x⌰
vp = V (2)
with h = height of the panel under consideration. Correspond-
ing to these panel displacements at the same initial point on
the framing, the resulting linearized deformations are

uf = 冉 冊
y ⫹ ȳ
H
UF (3)

vf = 0 (4)
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

with H = overall height of the wall. The above displacement


field for the sheathing and framing can be written more suc-
cinctly in terms of the global displacement vector DT = [UF ,

再冎 冋 册
Us, Ū, V̄, ⌰]

up 0 2y/h 1 0 ⫺y
vp 0 0 0 1 x
= D (5)
uf (y ⫹ ȳ)/H 0 0 0 0
vf 0 0 0 0 0
Extension of the above presentation to walls with multiple FIG. 3. Force-Displacement Response of Sheathing-to-Fram-
sheathing panels is straightforward. For the general case of a ing Connector under Monotonic and Cyclic Loading (Hysteretic
wall with Np sheathing panels, the number of degrees of free- Model Fitted to Connection Test Data for 50-mm-Long Spiral
Nail through 9.5-mm-Thick Oriented Strand Board Panel into
dom NDOF required to characterize the racking deformation
Spruce-Pine-Fir No. 2 Sawn Lumber Framing Member)
of the shear wall is NDOF = 4Np ⫹ 1.
The relative displacement between the sheathing and the
framing induces deformations in the sheathing-to-framing con- (producing the envelope curve OAI in Fig. 3) is modeled by
nectors within the wall. In Fig. 2(b), this deformation is iden- the following nonlinear load-deformation relationship
tified by ␦ for a generic connector fastened at one end to the F = sgn(␦) ⭈ (F0 ⫹ r1K 0兩␦兩) ⭈ [1 ⫺ exp(⫺K 0 兩␦兩/F0)], 兩␦兩 ⱕ 兩␦u 兩
framing member at point Q⬘ and at the other end to the sheath- (7a)
ing panel at point Q⬙. Evaluation of ␦ is given by
F = sgn(␦) ⭈ Fu ⫹ r2 K 0[␦ ⫺ sgn(␦) ⭈ ␦u], 兩␦u 兩 < 兩␦兩 ⱕ 兩␦F 兩 (7b)
␦ = 兹␦2u ⫹ ␦2v = 兹(up ⫺ uf)2 ⫹ (vp ⫺ vf)2 (6)
F = 0, 兩␦兩 > 兩␦F 兩 (7c)
As presented by (6), the connector deformation can be decom-
posed into a horizontal component ␦u and a vertical component This connector model, which was originally proposed by Fos-
␦v, which in turn can be determined from the global degrees chi (1977), is characterized by six physically identifiable pa-
of freedom for the wall through (5). rameters that must be fitted to experimental data: F0, K 0, r1,
Note that the kinematic assumptions outlined above for this r2, ␦u, and ␦F . Phenomenologically, (7) captures the crushing
shear wall model are essentially the same as those used pre- of the wood (framing and sheathing) along with yielding of
viously by Filiatrault (1990). the connector. Beyond the displacement ␦u, which corresponds
to the ultimate load Fu, the load-carrying capacity is reduced
Hysteretic Model of Sheathing-to-Framing because of connector withdrawal. Failure of the connector un-
Connectors der monotonic loading occurs at displacement ␦F .
Next, consider the load-deformation response of a connector
The load-deformation response of a dowel-type connector under the cyclic loading shown in Fig. 3. The basic path fol-
in a wood shear wall is highly nonlinear under monotonic lowing rules that define the proposed hysteretic model are
loading and exhibits pinched hysteretic behavior with strength identified and briefly discussed. In Fig. 3 load-displacement
and stiffness degradation under general cyclic loading (Dolan paths OA and CD follow the monotonic envelope curve as
and Madsen 1992a). Each connector behaves essentially as an expressed by (7). All other paths are assumed to exhibit a
elastoplastic pile (steel nail) embedded in a layered nonlinear linear relationship between force and deformation. Unloading
foundation (sheathing and framing material). Various research- off the envelope curve follows a path such as AB with stiffness
ers (Foschi 1974, 2000; Chui et al. 1998) have used this struc- r3 K 0. Here, both the connector and wood are unloading elas-
tural analogy to develop fairly sophisticated finite-element tically. Under continued unloading, the response moves onto
models for individual connectors. This approach is versatile path BC, which has reduced stiffness r4 K 0. Along this path,
and is capable of capturing the detailed cyclic response of a the connector loses partial contact with the surrounding wood
connector. However, it is computationally demanding to model because of permanent deformation that was produced by pre-
each connector within a shear wall in this manner. A simpler vious loading, along path OA in this case. The slack response
and more efficient approach is to develop a specific hysteretic along this path characterizes the pinched hysteresis displayed
model based on a minimum number of path-following rules by dowel connections under cyclic loading. Loading in the
that can reproduce the response of the connector under general opposite direction for the first time forces the response onto
cyclic loading. It is this latter method that is adopted for this the envelope curve CD. Unloading off this curve is assumed
study. elastic along path DE, followed by a pinched response along
Fig. 3 shows the assumed load-deformation behavior of a path EF, which passes through the zero-displacement intercept
connector under both monotonic loading and arbitrary cyclic FI , with slope r4 K 0. Continued reloading follows path FG with
loading. The response of a connector under monotonic loading degrading stiffness Kp, as given by
JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001 / 435

J. Struct. Eng., 2001, 127(4): 433-441


冉 冊

␦0 the wall response does not allow for a straightforward evalu-
Kp = K 0 (8) ation of (12) in terms of the global degrees of freedom. To
␦max
facilitate a solution, each sheathing-to-framing connector is
with ␦0 = (F0 /K 0); and ␣ = hysteretic model parameter that modeled using two uncoupled orthogonally oriented nonlinear
determines the degree of stiffness degradation. Note from (8) spring elements. This modeling approach has been utilized in
that Kp is a function of the previous loading history through many of the existing shear wall models discussed previously
the last unloading displacement ␦un off the envelope curve (Itani and Cheung 1984; Gupta and Kuo 1985; Dolan 1989;
(corresponding to point A in Fig. 3) so that White and Dolan 1995), even for the case of monotonic load-
ing where it is not actually required. Note that, under an im-
␦max = ␤␦un (9) posed deformation [as given by (6)], the use of two orthogonal
where ␤ = another hysteretic model parameter. The parameters uncoupled springs is only structurally equivalent, in terms of
␣ and ␤ are obtained by fitting the model to connection test resultant force and stiffness, to one spring if each spring is
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

data. A consequence of this stiffness degradation is that it also linear elastic with the same stiffness. For the hysteretic model
produces strength degradation in the response. If, on another of the sheathing-to-framing connector presented herein, the
cycle, the connector is displaced to ␦un, then the corresponding use of two uncoupled nonlinear springs will generally over-
force will be less than Fun that was previously achieved. This estimate the force and stiffness developed by the connector. A
strength degradation is shown in Fig. 3 by comparing the force method to adjust for the resulting stiffer wall response will be
levels obtained at points A and G. Also, with this model under outlined in a subsequent section.
continued cycling to the same displacement level, the force For racking of a shear wall, the external work is simply the
and energy dissipated per cycle stabilizes. This behavior is product of the applied top-of-wall framing displacement UF
close to what has been observed in connector tests (Dolan and and the induced top-of-wall lateral load FF . It then follows
Madsen 1992a), unless nail fatigue becomes a factor. that the right-hand side of (10) is given by
␦WE = FF ⭈ ␦UF = ␦DT F (13)
Governing Equations
T
where F = [FF , 0, 0, 0, 0] is the global force vector.
The equilibrium equations for the racking of a wood shear Substitution of (11)–(13) into (10) yields the nonlinear gov-
wall are obtained through application of the principle of virtual erning equilibrium equations for the racking response of the
displacements (PVD) shear wall assembly
␦WI = ␦(WF ⫹ Ws ⫹ WC)I = ␦(WS ⫹ WC)I = ␦WE (10) KS D = F (14)
where KS = KS (D) = global secant stiffness matrix, which is
a function of the displacement vector D. The components of
with the external work WI composed of contributions from the the global secant stiffness matrix can be evaluated explicitly
framing WF , sheathing WS , and connectors WC , and the exter- and are reported elsewhere (Folz and Filiatrault 2000). The
nal work WE arising from the applied racking load. As dis- formulation given above for obtaining the nonlinear equilib-
cussed previously, it is assumed in this model that the framing rium equations has been, for simplicity and clarity, presented
alone is a mechanism of pin-connected rigid members; hence, for a shear wall with only one sheathing panel. For this special
the internal work associated with this component of the wall case the model has only five degrees of freedom. Extension to
is zero. a wall with Np sheathing panels is straightforward and leads
The internal work absorbed during the uniform shear de- to a model with 4Np ⫹ 1 degrees of freedom to be solved.
formation of a sheathing panel is given by

冕 冉 冊 冉 冊
Displacement Control Solution Strategy
1 2Gbt 2Gbt
WS = ␶␥ dV = U 2S = DT BS BS D (11) The equilibrium equations, given by (14), are solved using
2 Vp h h an incremental-iterative displacement control solution strategy
where b, h, t, and Vp = width, height, thickness, and volume, (Batoz and Dhatt 1979; Ramm 1981). Suppose that the con-
respectively, of the panel under consideration. In the formu- figuration of the wall is known at load-step t and the solution
lation of (11), it is assumed the shear stress ␶ that develops in is sought at t ⫹ ⌬t. The incremental equilibrium equations can
the panel is linearly related to the strain field ␥ through the be written
panel’s shear modulus G. The final expression in (11) allows (t⫹⌬t)
KT ⌬D =
(t⫹⌬t) (t⫹⌬t)
⌬␭F0 ⫹ (t)R (15)
WS to be expressed in terms of the global displacement vector
D, by setting BS = [0, 1, 0, 0, 0]. with
The energy absorbed by all Nc sheathing-to-framing con-
nectors within a given sheathing panel is given by R = (t)F ⫺ (t)KS
(t) (t)
D (16)

冘 冉冕 冊
NC ␦j In (15), KT = global tangent stiffness matrix; ␭ = load factor
WC = Fj d␰ (12) applied to the reference global load vector F0; and R = global
j=1 0 residual force vector, which is rewritten in (16) in terms of
known quantities. To solve (15), the incremental global dis-
where Fj = connector force resulting from deformation of ␦j placement vector is decomposed into two parts
between the sheathing and the framing. Determination of Fj ,
for a given ␦j , is obtained from the rules that define the hys- (t⫹⌬t)
⌬D = ⌬␭⌬DI ⫹ (t)⌬DII
(t⫹⌬t)
(17)
teretic model presented previously. Using (5) and (6), (12) can
be expressed in terms of the global displacement vector D. which, in turn, produces two systems of equations to be solved
It has been observed that under monotonic loading of a (t⫹⌬t)
KT (t⫹⌬t)
⌬DI = F0; (t⫹⌬t)
KT (t⫹⌬t)
⌬DII = (t)R (18a,b)
shear wall, the deformation trajectory of each connector is es-
sentially unidirectional (Tuomi and McCutcheon 1978). Under Under displacement control, the top of wall translation of the
general cyclic loading of the wall, the displacement trajectory shear wall UF ⬅ D1 is prescribed. Hence, from (17) the fol-
of a connector can be bidirectional. This added complexity in lowing constraint equation is obtained:
436 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001

J. Struct. Eng., 2001, 127(4): 433-441


⌬D1 =
(t⫹⌬t)
⌬␭
(t⫹⌬t)
⌬D 1I ⫹
(t⫹⌬t)
⌬D1II = 0
(t⫹⌬t)
(19) computed sheathing-to-framing connector parameters obtained
from the experimental study (Durham 1998). Fig. 3 shows the
from which the increment in the load factor can be determined monotonic and cyclic load-displacement behavior of the
(t⫹⌬t)
⌬D 1II sheathing-to-framing connectors when the hysteretic connector
(t⫹⌬t)
⌬␭ = ⫺ (t⫹⌬t) (20) model is fitted with these parameters.
⌬D 1I The cyclic loading protocol used in this test program is
For a given displacement increment, Newton-Raphson itera- shown as an insert in Fig. 4. As is standard practice in shear
tions are performed on (18) until the new equilibrium config- wall testing, this cyclic loading protocol was scaled by the
uration of the shear wall is obtained to within a specified tol- wall’s load-displacement performance under monotonic load-
erance on the residual force vector. The total displacements ing. With this protocol, the wall is first subjected to three cy-
and forces acting on the shear wall are then updated from the cles with a maximum displacement corresponding to 50% of
previous increment the wall’s ultimate load established under monotonic testing.
This displacement level is denoted as ⌬0.5Pu in Fig. 4. The
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

D = (t)D ⫹ [(⌬t)⌬D I ⫹
(t⫹⌬t)
⌬D II];
(⌬t) (t⫹⌬t)
F= ␭F0
(t⫹⌬t)
(21a,b) protocol then has three cycles at ⌬0.8Pu, followed by one trail-
ing cycle back at ⌬0.5Pu and finishing with a unidirectional
The above presented solution strategy fails when the global push-over of the wall until failure.
tangent stiffness matrix KT is nonpositive definite. To over- The experimentally obtained load-displacement response of
come this limitation, an artificial linear spring (Ramm 1981) the shear wall under both monotonic loading and cyclic load-
is introduced at the top of the shear wall so that the combined ing is presented in Fig. 4. The corresponding predictions by
tangent stiffness matrix of the spring plus shear wall remains the numerical model, through the program CASHEW, are
positive definite over the cyclic loading protocol. This can be given in Fig. 5. A visual comparison of Figs. 4 and 5 shows
achieved by setting the axial stiffness of the artificial spring that good agreement is achieved between the model prediction
equal to the initial stiffness of the shear wall. The force de- and the experimental result for the prescribed cyclic loading.
veloped by the spring under the prescribed displacement UF is In particular, the observed stiffness and strength degradation
removed at the end of each loading step to obtain the top-of- that the test wall exhibited under the loading protocol was
wall force in the shear wall according to (21). fully captured by the numerical model. However, the model
It is well known that the hysteretic response of a shear wall did not perform as well in predicting the response during the
is largely determined by the hysteretic characteristics of the single trailing cycle of the loading protocol. Key results from
sheathing-to-framing connectors. In the model presented these two figures are the lateral load-carrying capacity Pu of
herein, the connector properties are defined in terms of a finite the wall and the corresponding drift ⌬u, which are summarized
set of path following rules. For the load-displacement response in Table 2. Under cyclic loading the difference between the
of the individual connectors to be captured, according to these test results and the model predictions are 7.8 and 9.1%, re-
rules, the step size in the displacement protocol for the overall spectively, for the ultimate load and displacement. Another
wall must not be too large. Unfortunately, this limiting step means of verifying the accuracy of the model is through the
size is not known a priori. To advance the solution, an adaptive
strategy has been implemented with a variable step size de- TABLE 1. Sheathing-to-Framing Connector Parameters for
termined by a bisectional search to ensure convergence at each 55-mm Long Spiral Nails (from Durham 1998)
step of the specified loading protocol.
The modeling features described herein for the cyclic re- K0 F0 FI ␦u
(kN/mm) r1 r2 r3 r4 (kN) (kN) (mm) ␣ ␤
sponse of wood shear walls have been incorporated in the
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
computer program CASHEW (Cyclic Analysis of SHEar
Walls) (Folz and Filiatrault 2000). 0.561 0.061 ⫺0.078 1.40 0.143 0.751 0.141 12.5 0.8 1.1

MODEL VERIFICATION
The predictive capabilities of the present numerical model
are compared with experimental results from full-scale mon-
otonic and cyclic shear wall tests performed under a separate
investigation (Durham 1998; Durham et al. 1999).
In this experimental study, the dimensions of the test shear
walls were 2.4 ⫻ 2.4 m. All framing material was 38 ⫻ 89
mm dimensional lumber. The top plate and end studs consisted
of double members, whereas the sole plate and the interior
studs were single members. Studs were spaced at 400 mm on
center. Conventional corner hold downs were used to prevent
overturning of the wall and to ensure a racking mode of de-
formation. The sheathing panels were 9.5-mm-thick oriented
strand board, with an assigned elastic shear modulus of 1.5
GPa. Three panels were used to sheath the wall: a 1.2 ⫻ 2.4
m panel covered the bottom half of the wall and two 1.2 ⫻
1.2 m panels covered the top half of the wall. Horizontal
blocking was used at midheight along the wall between the
sheathing panels. The sheathing-to-framing connectors were
pneumatically driven 50-mm-long spiral nails with a shank
diameter of 2.67 mm. Nails were spaced at 150 mm on center
along all panel edges and 300-mm spacing for all interior
studs. This testing program also collected sheathing-to-framing
connector data and fitted it to a hysteretic connector model FIG. 4. Experimental Monotonic and Cyclic Shear Wall Tests
similar to the one presented in this study. Table 1 lists the (Insert Shows Cyclic Testing Protocol)

JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001 / 437

J. Struct. Eng., 2001, 127(4): 433-441


Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

FIG. 6. Energy Absorbed during Cyclic Shear Wall Test: Ex-


FIG. 5. CASHEW Predictions of Monotonic and Cyclic Shear perimental Result and CASHEW Prediction
Wall Tests

TABLE 2. Summary of Test Results and Model Predictions un-


der Monotonic and Cyclic Loading
Monotonic
Loading Cyclic Loading
Pu ⌬u Pu ⌬u Ea
Parameter (kN) (mm) (kN) (mm) (kN-m)
(1) (2) (3) (4) (5) (6)
Test 17.4 57.4 20.4 66.0 2.59
Model 22.0 60.0 22.0 60.0 2.68
Difference (%) 26.4 4.5 7.8 9.1 3.5

evaluation of the energy absorbed by the wall under the cyclic


loading protocol


⌬F

Ea = FT ⌬ d⌬ (22)
0

In the evaluation of (22), which must be performed numeri-


cally, the limits of integration track the full displacement of
the wall under the loading protocol. The comparison of energy
absorbed during the cyclic test and the model’s prediction is
presented in Fig. 6, which reveals very good agreement. Nu-
merically, the difference in the total energy absorbed between FIG. 7. Monotonic Load-Displacement Response Using One-
the test result and the model prediction is only 3.5%. and Two-Spring Sheathing-to-Framing Connector Models in
Comparing the test result and the model prediction of the CASHEW
load-displacement response of the wall under monotonic load-
ing, as presented by Figs. 4 and 5 and as summarized in Table thogonal uncoupled nonlinear springs, as discussed previously.
2, reveals a fairly significant discrepancy, with the estimate of However, the specified connector spacing was adjusted so that
ultimate load differing by 26.4%. This difference may be at- the monotonic load-displacement response agreed, in terms of
tributable to the variability in construction quality between the energy absorbed by the wall up to a prescribed drift level, with
two test walls. Durham noted that with the wall tested mon- the prediction based on using only one nonlinear spring per
otonically there was ‘‘observed poor nailing’’ (Durham 1998). connector with the spacing unchanged. As shown in Fig. 7, if
This example reinforces the point that a purely deterministic this adjustment is not made, the model overpredicts the initial
evaluation should not be made between a single test result and wall stiffness and the ultimate load-carrying capacity. Surpris-
a model’s prediction of that test. It is to be expected that sim- ingly, this result has not been discussed in other research work
ilarly constructed shear walls will exhibit variability in their that has used two uncoupled nonlinear springs to model each
response under load. Unfortunately, the quantification of this connector.
variability has not been a primary objective of testing pro-
grams investigating shear wall behavior. MODEL APPLICATION
In the model verification study presented above, each As has been shown, the numerical model presented herein
sheathing-to-framing connector was represented by two or- is capable of predicting the load-displacement response of
438 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001

J. Struct. Eng., 2001, 127(4): 433-441


wood shear walls under cyclic loading. To perform this anal- havior, strength and stiffness degradation, etc.) as those exhib-
ysis for a given shear wall configuration, only the shear mod- ited by an individual sheathing-to-framing connector under cy-
ulus of the sheathing panels and cyclic test data from the clic loading (Doland and Madsen 1992b). This observation can
sheathing-to-framing connectors are required as input data. be made from this study as well by comparing the cyclic re-
It is obvious that more information can be obtained about sponse shown in Figs. 3 and 4. Consequently, the hysteretic
the structural performance of a shear wall through performing model presented earlier, which was applied to sheathing-to-
a cyclic analysis compared to just a monotonic push-over anal- framing connectors, can be used to represent the global re-
ysis. It is equally obvious that the dynamic analysis of a shear sponse of the shear wall under cyclic loading with appropri-
wall provides still greater information, especially as it relates ately applied parameter values. These hysteretic model
to the seismic design problem. In a number of previous stud- parameters for the wall can be identified through a nonlinear
ies, the dynamic behavior of shear walls has been investigated functional minimization procedure. The objective function to
using nonlinear SDOF models [e.g., Stewart (1987) and Fol- be minimized is the cumulative error between the restoring
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

iente (1995)]. In these studies, hysteretic elements, which force FF developed in the shear wall as predicted by the SDOF
modeled the global wall behavior under cyclic loading, were hysteretic model and that obtained from the CASHEW model
fitted to experimental data obtained from full-scale shear wall when subjected to the same cyclic loading protocol.
tests. These hysteretic elements were incorporated into nonlin- The procedure outlined above will now be applied to the
ear SDOF dynamic analysis programs to numerically predict 2.4 ⫻ 2.4 m shear wall considered earlier in this study. As
the seismic response of shear walls. This research work has part of the experimental study by Durham (1998), this shear
shown that acceptable results can be obtained through a SDOF wall configuration was tested dynamically on a shake table.
idealization of a wood shear wall. These previously cited stud- The wall carried a seismic mass of 5,450 kg and was subjected
ies, however, were limited by the need of full-scale cyclic test to the E-W component of the 1992 California Landers earth-
data to calibrate the hysteretic shear wall elements. The same quake, recorded at Joshua Tree Station, scaled to have a peak
procedure is adopted herein with the exception that the ground acceleration of 0.36g. This particular earthquake record
CASHEW model will be used to calibrate the SDOF hysteretic has the distinguishing feature of two separated segments of
model for the dynamic analysis. In this way the dependency high amplitude acceleration. The shake table’s reproduction of
on full-scale shear wall test data is eliminated. this record, as a base acceleration input to the test shear wall,
The idealization of a shear wall as an equivalent SDOF is shown in Fig. 9(a). The response of the test wall to this
system, subjected to seismic loading, is illustrated in Fig. 8. acceleration record, in terms of relative displacement time his-
The hysteretic behavior of the wall is captured by a nonlinear tory, is presented in Fig. 9(b). As seen from Fig. 9(b), two
spring having a force-displacement relationship FF [UF (t)], separated segments of high amplitude drift are produced in the
with FF as the top of wall force, or base shear, corresponding test wall by this earthquake record.
to a top of wall displacement UF . The equation of motion for To obtain the SDOF response prediction for this wall, using
this SDOF system can be expressed (23), it is necessary to first determine the parameters of the
¨ F (t) ⫹ cU
˙ F (t) ⫹ FF [UF (t)] = ⫺mag(t) hysteretic element and the viscous damping coefficient. Using
mU (23)
where m = seismic mass carried by the wall; c = linear viscous
damping coefficient; U̇F (t) and ÜF (t) = velocity and acceler-
ation, respectively, of the top of the wall relative to the moving
base; and ag(t) = ground acceleration. The solution of (23) is
achieved using the Newmark constant average-acceleration
time integration scheme. An energy balance calculation is per-
formed to ensure accuracy of the solution.
It is well known that the hysteretic response of a typical
shear wall has the same defining characteristics (pinched be-

FIG. 9. Seismic Analysis of Shear Wall: (a) Shake Table Accel-


eration Record (Landers Earthquake Peak Ground Acceleration
ⴝ 0.36g); (b) Experimental Displacement Time History; (c)
FIG. 8. Modeling Shear Wall as SDOF System SDOF Model Prediction

JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001 / 439

J. Struct. Eng., 2001, 127(4): 433-441


TABLE 3. Fitted Hysteretic Parameters for SDOF Shear Wall hysteretic response predicted by the SDOF model is fully con-
Model tained within the monotonic envelope curve obtained by the
K0 F0 FI ⌬u CASHEW model. This confirms the accuracy of the calibra-
(kN/mm) r1 r2 r3 r4 (kN) (kN) (mm) ␣ ␤ tion procedure that was applied to the SDOF model.
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
CONCLUSIONS
1.44 0.081 ⫺0.022 1.31 0.064 15.1 3.13 60.0 0.74 1.10
A simple numerical formulation for the structural analysis
of wood-framed shear walls under arbitrary cyclic loading has
TABLE 4. Summary of Test Results and Model Predictions un- been elaborated based on the hysteretic properties of sheath-
der Dynamic Loading ing-to-framing connectors. The resulting numerical model, in-
⌬1 ⌬2 ⌬3 ⌬4 corporated in the computer program CASHEW, is able to pre-
Parameter [mm (s)] [mm (s)] [mm (s)] [mm (s)] dict the load-displacement response and energy dissipation
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

(1) (2) (3) (4) (5) characteristics of wood shear walls, with or without opening,
Test 37.3 (9.67) ⫺31.9 (10.17) ⫺60.0 (26.4) 51.7 (26.9) under arbitrary quasi-static cyclic loading.
Model 39.4 (9.55) ⫺33.5 (9.95) ⫺45.0 (26.3) 54.1 (26.7) The model has been verified against full-scale tests of wood-
Difference (%) 5.6 5.0 25.0 4.44 framed shear walls subjected to monotonic and cyclic loading.
The predictions of the model agreed well with the experi-
mental results.
As an application of the proposed model, the parameters of
a generalized nonlinear SDOF wall system have been cali-
brated based on the cyclic shear wall response predicted by
the CASHEW model. This SDOF system can then be used to
obtain the seismic response of a shear wall under arbitrary
earthquake ground motions. Comparisons of the numerical
predictions with the results of shake table tests conducted on
a shear wall with a lumped mass at its top confirmed the ad-
equacy of this SDOF approach.
As an extension of this study, the SDOF calibration proce-
dure can be generalized and used for the cyclic and seismic
analyses of complete 3D wood-framed buildings. The contri-
bution of each shear wall to the global lateral load resisting
system can be represented by an equivalent SDOF hysteretic
element, calibrated through the CASHEW model, using cyclic
test data for the sheathing-to-framing connectors. The writers
are currently involved in the development of this approach.
ACKNOWLEDGMENTS
The work presented herein was carried out under contract to the
consortium of Universities for Research in Earthquake Engineering
(CUREe), Richmond, Calif., as part of the CUREe-Caltech WoodFrame
Project (‘‘Earthquake hazard mitigation of woodframe construction’’)
under a grant administered by the California Office of Emergency Ser-
FIG. 10. SDOF Model Prediction of Hysteretic Response of vices, Mather Field, Calif., and funded by the Federal Emergency Man-
Shear Wall under Seismic Loading and Monotonic Envelope agement Agency, Washington, D.C. The financial support provided by
Curve Predicted by CASHEW these funding agencies is gratefully acknowledged. The writers thank Dr.
Helmut Prion of the University of British Columbia, Vancouver, for gra-
ciously providing the shear wall test data from Jennifer Durham’s Mas-
the load-displacement response predicted by the CASHEW ter’s thesis.
model, as shown in Fig. 5, the system identification procedure
presented above was utilized to obtain the defining parameters APPENDIX. REFERENCES
for the SDOF hysteretic element. The results of this analysis Batoz, J. L., and Dhatt, G. (1979). ‘‘Incremental displacement algorithms
are summarized in Table 3. Because the hysteretic element has for nonlinear problems.’’ Int. J. Numer. Methods in Engrg., 14, 1262–
been fitted to capture the global response of the shear wall, 1267.
with the various energy dissipating mechanisms of the wall Chui, Y. H., Ni, C., and Jaing, L. (1998). ‘‘Finite-element model for nailed
wood joints under reversed cyclic load.’’ J. Struct. Engrg., ASCE,
contributing to this response, the viscous damping coefficient 124(1), 96–103.
c in (23) is assigned a nominal value of only 1% of critical. Commins, A., and Gregg, R. C. (1994). ‘‘Effect of hold-downs and stud-
This is consistent with what has been suggested by other re- frame systems on the cyclic behavior of wood shear walls.’’ Proc., Res.
search work (Foliente 1995). Needs Workshop on Anal., Des. and Testing of Timber Struct. under
The predicted relative displacement time history of the Seismic Loads, G. C. Foliente, ed., Forest Products Lab., University of
SDOF shear wall model is shown in Fig. 9(c). A visual com- California, Richmond, Calif., 142–146.
Dolan, J. D. (1989). ‘‘The dynamic response of timber shear walls.’’ PhD
parison of Figs. 9(b and c) shows that the SDOF model pro- thesis, University of British Columbia, Vancouver.
duces reasonably good correlation with the test result for the Dolan, J. D., and Foschi, R. O. (1991). ‘‘Structural analysis model for
relative displacement time history. Further numerical compar- static loads on timber shear walls.’’ J. Struct. Engrg., ASCE, 117(3),
isons between the test results and the model predictions are 851–861.
given in Table 4. Dolan, J. D., and Madsen, B. (1992a). ‘‘Monotonic and cyclic nail con-
For completeness, Fig. 10 shows the SDOF model predic- nection tests.’’ Can. J. Civ. Engrg., Ottawa, 19(1), 97–104.
Dolan, J. D., and Madsen, B. (1992b). ‘‘Monotonic and cyclic tests of
tion of the hysteretic response of the shear wall resulting from timber shear walls.’’ Can. J. Civ. Engrg., Ottawa, 19(4), 415–422.
the seismic acceleration record. Also plotted in Fig. 10 is the Durham, J. P. (1998). ‘‘Seismic response of wood shearwalls with over-
CASHEW prediction of the monotonic load-displacement re- sized oriented strand board panels.’’ MASc thesis, University of British
sponse of the test wall. As can be seen from this figure, the Columbia, Vancouver.

440 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001

J. Struct. Eng., 2001, 127(4): 433-441


Durham, J., Prion, H. G. L., Lam, F., and He, M. (1999). ‘‘Earthquake —Joints made with mechanical fasteners—Quasi-static reversed-cyclic
resistance of shearwalls with oversize sheathing panels.’’ Proc., 8th test method.’’ Draft Document ISO TC 165/SC N., Geneva.
Can. Conf. on Earthquake Engrg., Canadian Association for Earth- Itani, R. Y., and Cheung, C. K. (1984). ‘‘Nonlinear analysis of sheathed
quake Engineering, Dept. of Civil Engrg., University of B.C., Vancou- wood diaphragms.’’ J. Struct. Engrg., ASCE, 110(9), 2137–2147.
ver, Canada, 161–166. Karacabeyli, E., and Ceccotti, A. (1996). ‘‘Test results on the lateral re-
European Committee for Standardization (CEN). (1995). ‘‘Timber struc- sistance of nailed shear walls.’’ Proc., Int. Wood Engrg. Conf., Vol. 2,
tures—Test methods-cyclic testing of joints made with mechanical fas- Omnipress, Madison, Wis., 179–186.
teners.’’ pr EN 12512, Brussels. Krawinkler, H., Parisi, F., Ibarra, L., Ayoub, A., and Medina, R. (2000).
Filiatrault, A. (1990). ‘‘Static and dynamic analysis of timber shear ‘‘Development of a testing protocol for wood frame structures.’’ CUREe
walls.’’ Can. J. Civ. Engrg., Ottawa, 17(4), 643–651. Publ. No. W-02, Consortium of Universities for Research in Earthquake
Foliente, G. C. (1995). ‘‘Hysteresis modeling of wood joints and struc- Engineering, Richmond, Calif.
tural systems.’’ J. Struct. Engrg., ASCE, 121(6), 1013–1022. Lam, F., Prion, H. G. L., and He, M. (1997). ‘‘Lateral resistance of wood
Folz, B., and Filiatrault, A. (2000). ‘‘CASHEW—Version 1.0: A com- shear walls with large sheathing panels.’’ J. Struct. Engrg., ASCE,
puter program for cyclic analysis of wood shear walls.’’ Rep. No. SSRP- 123(12), 1666–1673.
Downloaded from ascelibrary.org by Politehnica University of Timisoara on 05/22/17. Copyright ASCE. For personal use only; all rights reserved.

2000/10, Struct. Sys. Res. Proj., Dept. of Struct. Engrg., University of Ramm, E. (1981). ‘‘Strategies for tracing the nonlinear response near limit
California-San Diego, La Jolla, Calif. points.’’ Nonlinear finite element analysis in structural mechanics, W.
Foschi, R. O. (1974). ‘‘Load-slip characteristics of nails.’’ Wood Sci., 7(1), Wunderlich, E. Stein, and K.-J. Bathe, eds., Springer-Verlag, Berlin, 63–
69–74. 89.
Foschi, R. O. (1977). ‘‘Analysis of wood diaphragms and trusses, Part 1: Rose, J. D. (1994). ‘‘Performance of wood structural panel shear walls
Diaphragms.’’ Can. J. Civ. Engrg., Ottawa, 4(3), 345–362. under cyclic (reversed) loading.’’ Proc., Res. Needs Workshop on Anal.,
Des. and Testing of Timber Struct. under Seismic Loads, G. C. Foliente,
Foschi, R. O. (2000). ‘‘Modeling the hysteretic response of mechanical
ed., Forest Products Lab., University of California, Richmond, Calif.,
connections for wood structures.’’ Proc., World Conf. on Timber
129–141.
Engrg., Dept. of Civ. Engrg., Dept. of Wood Science and School of
SEAOSC. (1997). ‘‘Standard method of cyclic (reversed) load test for
Architecture, University of B.C., Vancouver, Canada.
shear resistance of framed walls for buildings.’’ Structural Engineers
Gupta, A. K., and Kuo, G. P. (1987). ‘‘Wood-framed shear walls with
Association of Southern California (SEAOSC) Report, Whittier, Calif.
uplifting.’’ J. Struct. Engrg., ASCE, 113(2), 241–259. Shenton, H. W., Dinehart, D. W., and Elliott, T. E. (1998). ‘‘Stiffness and
Gupta, A. K., and Kuo, P.-H. (1985). ‘‘Behavior of wood-framed shear energy degradation of wood frame shear walls.’’ Can. J. Civ. Engrg.,
walls.’’ J. Struct. Engrg., ASCE, 111(8), 1722–1733. Ottawa, 25(3), 412–423.
Gutkowski, R. M., and Castillo, A. L. (1988). ‘‘Single- and double- Skaggs, T. D., and Rose, J. D. (1996). ‘‘Cyclic load testing of wood struc-
sheathed wood shear wall study.’’ J. Struct. Engrg., ASCE, 114(6), tural panel shear walls.’’ Proc., Int. Wood Engrg. Conf., Vol. 2, Omni-
1268–1284. press, Madison, Wis., 195–200.
He, M., Lam, F., and Prion, H. G. L. (1998). ‘‘Influence of cyclic test Stewart, W. G. (1987). ‘‘The seismic design of plywood sheathed shear
protocols on performance of wood-based shear walls.’’ Can. J. Civ. walls.’’ PhD thesis, University of Canterbury, Christchurch, New Zea-
Engrg., Ottawa, 25(3), 539–550. land.
He, M., Magnusson, H., Lam, F., and Prion, H. G. L. (1999). ‘‘Cyclic Tuomi, R. L., and McCutcheon, W. J. (1978). ‘‘Racking strength of light-
performance of perforated wood shear walls with oversize OSB pan- frame nailed walls.’’ J. Struct. Engrg., ASCE, 104(7), 1131–1140.
els.’’ J. Struct. Engrg., ASCE, 125(1), 10–18. White, M. W., and Dolan, J. D. (1995). ‘‘Nonlinear shear-wall analysis.’’
International Standards Organization (ISO). (1999). ‘‘Timber structures J. Struct. Engrg., ASCE, 121(11), 1629–1635.

JOURNAL OF STRUCTURAL ENGINEERING / APRIL 2001 / 441

J. Struct. Eng., 2001, 127(4): 433-441

You might also like