Anda di halaman 1dari 14

Journal of Constructional Steel Research 64 (2008) 87–100

www.elsevier.com/locate/jcsr

Sources of elastic deformation in steel frame and framed-tube structures:


Part 1: Simplified subassemblage models
Finley A. Charney ∗ , Rakesh Pathak
Department of Civil and Environmental Engineering, Virginia Tech, Blacksburg, VA 24061, United States

Received 4 August 2006; accepted 8 May 2007

Abstract

Three simple models for including the effect of beam–column joint deformation in the analysis of steel moment resisting frame and framed
tube structures are presented. The first model, called the Fictitious Joint model, is based on two-dimensional frame analysis and is useful for
preliminary analysis only. The second model, called the Krawinkler Joint model, and the third model, known as the Scissors Joint model, use an
assembly of rigid links and rotational springs to represent the joint, and may be used in preliminary and final analysis of full structural systems.
All derivations are provided in the form of “displacement participation factors”, which allow a detailed breakdown of the various components of
subassemblage displacement.
When applied to isolated beam–column subassemblages, it is shown that all three modeling approaches produce the same general expression
for computing deflections arising from shear deformations in the panel zone region. However, the Krawinkler and Scissors models do not include
the effects of flexural deformation within the beam–column joint, whereas the Fictitious Joint model does. While not the dominant source of
deformation, it is shown in the paper that the effects of flexural deformations in the beam–column joint should not be ignored.
It is also shown in this paper that the overall displacements predicted by the simplified models correlate very well with displacements computed
from detailed three dimensional finite element analysis of the same subassemblage. However, the finite element analysis approach, taken alone,
is not capable of providing a breakdown of the subassemblage displacements into components, such as panel zone shear, or column joint flexure.
Part 2 of the paper presents a method for providing this information from the results of detailed finite element analysis.
c 2007 Elsevier Ltd. All rights reserved.

Keywords: Steel structures; Moment resisting frames; Panel zone deformations; Structural analysis

1. Introduction and background Krawinkler [5] was the first to propose a mechanical
beam–column joint model, which is applicable to both
The effect of beam–column joint deformations on the lat- linear and nonlinear analysis. A similar, but less complex
eral load response of steel moment resisting frames and framed model, called the Scissors model has been proposed by
tubes has been studied in detail [1–16]. In these references, a variety of researchers. As documented by Charney and
there is general agreement that it is essential that beam–column Marshall [16], there appears to be confusion as to the
joint deformations be included in structural analysis, and vari- properties to use in the mechanical models, particularly in the
ous approaches are provided for including such effects. These Scissors model. However, when properly used, the mechanical
approaches may basically be broken into two types. In the first models are reasonably accurate when compared to experimental
type, the beam–column joint region is modeled explicitly, us- results [12]. One of the key advantages of the mechanical
ing an assemblage of rigid links, rotational springs, and shear models is that they may be used in the structural analysis of
panels. The second approach uses standard frame analysis, with full structural systems.
joint deformations being considered through the use of a mod- In the standard frame analysis approach, the deformations
ified force distribution in the beam–column joint region. in the joint region are considered through a modification
of the forces in the joint regions of the structure. This
∗ Corresponding address: Department of Civil and Environmental Engineer-
technique is most often applied to isolated subassemblages,
ing, 200 Patton Hall, Mail Stop 105D, Blacksburg, VA 24061, United States.
Tel.:+1 540 231 1444; fax: +1 540 231 7532. and is not applicable to full frames unless certain compatibility
E-mail address: fcharney@vt.edu (F.A. Charney). requirements are relaxed. For example, the method may be used

c 2007 Elsevier Ltd. All rights reserved.


0143-974X/$ - see front matter
doi:10.1016/j.jcsr.2007.05.008
88 F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100

if a two-dimensional frame is represented as an assemblage of Table 1


subassemblages, where compatibility is relaxed by assuming Basic subassemblages considered for analysis
that there are points of inflection at the mid-span of girders and Designation Beam size Beam span Column size Column
mid-height of columns. The standard frame analysis approach (m (ft)) height
may also be used to estimate the effect of beam–column joint (m (ft))
deformations on a structure which has been analyzed without A W770 × 220 4.57 (15) W360 × 382 3.81(12.5)
explicit modeling of the joints. In this case, a virtual work based (W30 × 148) 6.10 (20) (W14 × 257)
post-processor is used, as described in Charney [8]. 7.62 (25)
There is one significant difference between the mechanical
B W770 × 220 4.57 (15) W610 × 372 3.81(12.5)
models and the frame models, and this has to do with the
(W30 × 148) 6.10 (20) (W24 × 250)
modeling of flexural deformations in the beam–column joint. 7.62 (25)
In the mechanical models, deflection resulting from flexural
flexibility in the joint is ignored, and stiffness associated with
bending in the column flanges is included. In the frame model, on (gross) simplifications of the pattern of shear and flexural
exactly the opposite is done; the bending in the joint region is stresses within the joint region.
considered, and added stiffness associated with column flange The three different joint models are presented through the
bending is ignored. The result is that a given subassemblage use of a virtual work procedure, called the Displacement
modeled with the two different approaches will have different Participation Factor (DPF) approach [8,18,19]. After the
stiffness, with the mechanical model being the stiffer of the two. theoretical background is presented, the DPF procedure is used
Because the effect of flexural deformations in the joint to analyze two basic subassemblages. Each is run with three
region is significant [9], it is desirable to modify the mechanical different beam spans (4.57 m (15 ft), 6.10 m (20 ft), and
models to include such deformations. Downs [12] provided 7.62 m (25 ft)), and with and without beam flange continuity
preliminary recommendations incorporating such deformations plates. The basic list of subassemblages considered is shown in
in the Krawinkler model, but final implementation of the Table 1. Individual subassemblages are designated by a three
approach requires a better understanding of the true pattern part symbol consisting of basic designation, beam span (in
and effect of deformations within the beam–column joint. feet), and presence of continuity plates. For example, B20c
This understanding is obtained through detailed finite element represents basic subassemblage B with a 20 ft. (6.10 m) beam
analysis, and is presented in Part 2 of this paper [17]. span, with continuity plates. The same subassemblage without
continuity plates is B20n.
2. Objectives and scope When continuity plates are used, they are positioned top and
bottom, on each side of the column web, and are of the same
The objective of the work presented in this paper is to present thickness as the flange of the beam. The width of the continuity
and determine the overall accuracy of three simple models for plate is taken as the lesser of the width of the projecting
computing the effect of beam–column joint deformations on flange width of the beam or column. Web doubler plates are
lateral load induced displacements of steel moment resisting not considered in the analysis performed herein, but they are
frames and framed tube structures. The accuracy is assessed included in all derivations. Even though the doubler plates are
by comparing the total subassemblage drift computed by the not explicitly included in the examples, the benefits of using
simplified models with results obtain from detailed three- doubler plates to reduce drift are briefly discussed.
dimensional finite element analysis. After the derivations are presented, the results of the
It is important to note that this paper does not specifically analyses of the subassemblages indicated in Table 1 are
address the influence of axial deformations in the columns. This presented and discussed.
can be a very significant source of deformation in tall steel
frames, and particularly in framed-tubes. An earlier conference 3. Overview of displacement participation factor (DPF)
paper by Charney [8] does address axial deformations in frames approach
and tube structures as tall as 50 stories. Axial deformations are
included in a similar manner in part 3 of the current paper. Part Consider the structure of Fig. 1(a), which is subjected to the
3 will be published in a different issue of the same journal in “Real” lateral loads shown on the left side of the structure.
which parts 1 and 2 appeared. A simple two-dimensional frame analysis program has been
The first two simplified models are mechanical models, and used to determine the lateral (x) deflection at the upper right of
the third is a frame model. The mechanical models are referred the structure (node 28). This deflection is designated as ∆28x.
herein as the Krawinkler Joint (KJ) model and the Scissors The analysis was based on centerline dimensions, with axial,
Joint (SJ) model. A detailed history and description of the flexural, and shear deformations included.
mechanical models is provided in Charney and Marshall [16]. It is now desired to determine how much the highlighted
The frame model is called the Fictitious Joint (FJ) model. The column (C10) contributes to the designated deflection. To
name for this model comes from the fact that the beam–column do so, the real loads are removed, and a virtual force Q̂ is
joint is not modeled explicitly. Instead, it is represented as a part applied at the location of and in the direction of ∆28x. The
of the column, and the deformation in the joint region is based column’s contribution to the displacement, called the column’s
F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100 89

Fig. 1. Planar frame under real or virtual loading.

Displacement Participation Factor, is given by the familiar


virtual work expression
"Z
1 H P(x) P̂(x) Z H
M(x) M̂(x)
∆28x
∇C10,T = dx +
Q̂ 0 E AC 0 E IC
#
V (x)V̂ (x)
Z H
+ dx (1)
0 G AC S

where P(x), V (x) and M(x) are the “real” axial force,
shear, and moment functions along the length of the element,
P̂(x), V̂ (x) and M̂(x) are the “virtual” force functions in the
element (arising from the application of Q̂), H is the length
of the element, E and G are the elastic and shear modulus,
respectively, and AC , IC and AC S , are the cross sectional
properties of the column.
Fig. 2. Typical beam–column subassemblage.
It is clear from Eq. (1) that the column’s contribution to the
deflection consists of axial, flexural, and shear components. If A similar approach could be used to determine how the
DPFs for all of the beams and columns of the structure were highlighted subassemblage (Sub. A) of Fig. 1(a) contributes
available, it would be seen that to a global displacement, such as the x-direction displacement
at node 28. In this case the contribution of the beam, column,
ncols nbeams
X
∆28x
X and beam–column joint region would be computed. A different
∆28x = ∇Ci,T + ∇ B∆j,T
28x
(2)
i=1 j=1
virtual load pattern, such as that shown in Fig. 1(b), could
be used to determine subassemblage A’s contribution to the
where the summation ranges over the total number of columns interstory drift (δ3) at level three of the structure.
and beams. It is also of interest to analyze an isolated subassemblage
As seen in Eq. (1), the basic symbol for a DPF is the inverted that is not part of a larger structure. Such a subassemblage
triangle, with a double subscript and a single superscript, (which might be a laboratory specimen) is shown in Fig. 2.
∆ . The superscript represents the “designated
as follows: ∇C,S Here, DPFs are used to determine how the column, beam,
displacement”, which in the case of the previous example, was and joint of the subassemblage contribute to the total drift,
∆28x. The first subscript, C, represents the named component δ, as illustrated in Fig. 3. This total drift is the designated
of the structure for which the DPF is computed. There are displacement for the subassemblage, and hence, the DPF
numerous possibilities here, but B, C, and J serve as examples terms for isolated subassemblage analysis will have δ as a
for beam, column, and joint, respectively. The subscript, S, superscript. All derivations presented in this paper are based
represents the deformation source, and can be either A, F, S, or on the assumption that the subassemblage is isolated.
T for axial, flexure, shear, or total, respectively. For example, In the remainder of this paper, derivations are provided for
from Eq. (1) it is seen that computing DPFs for the FJ, KJ, and SJ models. When doing
so, the subassemblage is divided into column, beam, and joint
∆28x ∆28x ∆28x ∆28x
∇C10,T = ∇C10,A + ∇C10,F + ∇C10,S . (3) components. DPFs for the portions of the column and beam
90 F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100

It is desired to determine how much the various components


of the subassemblage contribute to the drift along the height of
the subassemblage, so equal and opposite virtual forces Q̂ are
applied at the top and bottom of the column, as shown in Fig. 3.
The shear in the column resulting from this force is designated
as V̂C , and all other virtual subassemblage forces are related
to this force. A carat (ˆ) over the appropriate symbol identifies
quantities as virtual. While the derivation would be simplified
by using a unit value for Q̂ this is not done because the presence
of Q̂ in the various equations provides force unit consistency.
Maintaining Q̂ in the derivations also allows the formulas to be
used to determine a subassemblage’s contribution to a global
displacement as shown in Fig. 1. In this case, VC is taken as the
average of the real shears in the upper and lower half column,
and V̂C is the average of the virtual shears in the upper and
Fig. 3. Beam–column subassemblage used in analysis. lower half column. Note, however, that an additional source
of deformation, due to beam and column axial deformations,
outside the joint region are identical for each of the models, would need to be included.
and these are called the clear span DPF components. The DPF The components analyzed consist of the clear span region
for the joint region represents the portion of the column which of the column, the clear span region of the beam, and the
passes through the joint, the beam flange continuity plates, and portion of the column that is described as the joint. The joint
the doubler plates. includes beam flange continuity plates and doubler plates,
After the simple model derivations are presented, the results when present. Only shear and flexural deformations need to
of the analysis of subassemblage A20c and A20n are described be considered because there are no axial forces acting on the
in detail. The purpose of this analysis and related discussion is isolated subassemblage.
to describe basic features of the behavior of the subassemblage, The total displacement along the height of the subassem-
and to set up the discussion for the detailed finite element blage is the same as the total DPF for the subassemblage, as
analysis. follows:
Finally, comparisons for the simplified and detailed analysis
of all of the subassemblages are presented and discussed. A δ = ∇Tδ = ∇C,F
δ δ
+ ∇C,S δ
+ ∇ B,F δ
+ ∇ B,S + ∇ δJ,F + ∇ δJ,S . (6)
general statement on the accuracy of the analysis is made, and
recommendations are provided for use by the design profession. The shears and moments in the clear spans of the beam
and column are easily determined because the subassemblage
4. Computation of DPF for the fictitious panel zone model is statically determinate. The forces inside the joint region
are indeterminate, however, and a simplifying assumption is
The subassemblage to be analyzed is shown in Fig. 2. This required to resolve these forces. The assumption made is
subassemblage is extracted from a frame by assuming that that the beam and column moments that enter the joint are
inflection points occur at mid span of the beams on either side of completely resolved into flange force couples. This is shown in
the joint, and at mid-height of the columns above and below the the free-body diagram of Fig. 4, where the subassemblage has
joint. It is further assumed that the beams on either side of the been divided into the component parts. Note that the lower half-
joint are of the same section and length, and that the columns column and the left half-beam have been omitted from Fig. 4 for
above and below the joint are the same section and length (but clarity.
not necessarily the same as the beam). The flange forces are designated as FBF for the beams and
To simplify the derivations, the dimensionless parameters α FCF for the columns, and are computed from equilibrium as
and β are defined: follows:

dCe 0.5L(1 − α)(VC H/L) 0.5VC (1 − α)


α= (4) FBF = = (7)
L βH β
d Be 0.5VC H (1 − β)
β= (5) FCF = . (8)
H αL
where L is the length of the beam (the bay width) and H is Force, shear, and moment diagrams for the full height of the
the height of the column (story height), and dCe and d Be are column are shown in Fig. 5. Similar diagrams for the beam are
the effective depths of the column and beam, respectively. As presented in Fig. 6. The force diagram for the column shows
shown in Fig. 2, these depths are taken as the distance between the left and right beam flange forces entering the column at
the centers of the flanges of the sections. Another possible the top and bottom of the joint region. Using Fig. 5, it is seen
definition of the effective depth, using out-to-out dimensions, that the horizontal shear in the joint region of the column is
is discussed later. VCJ = VC − 2FBF , which simplifies to
F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100 91

(1 − α − β)
VCJ = VC . (9)
β
Note that the shear force is taken as a positive quantity.
The vertical shear force, VBJ , in the joint region of the column
is similarly determined. From Fig. 6, it is seen that VBJ =
VB − 2FCF , which after some substitution and manipulation
results in
VC H (1 − α − β)
VBJ = . (10)
L α
It is important to note that the shear stress in the joint
region is the same whether computed from the horizontal or
the vertical shear force:
VCJ VBJ VC (1 − α − β)
τJ = = = (11a)
αLt p β Htp αβ Lt p
where t p is the thickness of the panel zone including doubler
plates, if present. If the numerator and denominator of Eq. (11a)
Fig. 4. Free body diagram of subassemblage components. are multiplied by H , the expression for shear stress may be
simplified to
VC H (1 − α − β) VC H (1 − α − β)
τJ = = (11b)
αβ L H t p vp
where v p = αβ L H t p is the effective volume of the panel zone.
From the moment diagrams of Figs. 5 and 6, it is seen that
the moments at the center of the column and beam, MCC and
MBC , respectively, are not exactly zero. The column moment is
MCC = 0.5H VC − 0.5β H (2FBF )
which simplifies to
MCC = 0.5VC H α. (12)
Similarly, the moment at the center of the beam is
VC H
MBC = 0.5L − 0.5αL(2FCF )
L
which simplifies to
Fig. 5. Column force diagrams.
MBC = 0.5VC Hβ. (13)
Clearly, equilibrium is not exactly satisfied at the center of
the joint because MCC and MBC are not equal. This small error
is due to the assumption that all of the moment is resisted by
the flange of the beam and column.
Given the moments and forces along the length of the
elements, it is now possible to derive the displacement
participation factors. This will be done first for the clear span
portions of the beam and column, and then for the joint region.
The real shears and moments in the clear span region of
the subassemblage are shown in Figs. 5 and 6. The virtual
shears and moments are identical, with the exception that (for
example) V̂C is used in lieu of VC . Using these forces, the DPFs
for the portion of the beams and columns outside the joint are
as follows:
2 0.5H (1−β) VC V̂C V̂C VC H (1 − β)
Z
δ
Fig. 6. Beam force diagrams. ∇C,S = dx = (14)
Q̂ 0 G A SC Q̂ G A SC
92 F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100

2
Z 0.5L(1−α) VC H V̂C H V̂C VC H 2 (1 − α)
δ L L
∇ B,S = dx = (15)
Q̂ 0 G AS B Q̂ LG A S B
2
Z 0.5H (1−β)
VC x V̂C x
δ
∇C,F = dx
Q̂ 0 E IC
V̂C VC H 3 (1 − β)3
= (16)
Q̂ 12E IC
V H x V̂C H x
2 0.5L(1−α) CL
Z
δ L
∇ B,F = dx
Q̂ 0 E I B
V̂C VC H 2 L(1 − α)3
= . (17)
Q̂ 12E I B
The terms A SC and A S B represent the effective shear areas
of the column and beam cross sections. For all computations
presented in this paper, these areas are computed using Fig. 7. Effective section for beam joint flexure.
Cowper’s detailed formula [20]. A complete summary of the
computation of shear areas in wide flange sections is provided
in Charney et al. [21].
The DPFs as presented in Eqs. (14) through (17) include
columns and beams on both sides of the joint (hence the “2”
on the left of the integral sign). DPFs for a centerline analysis
of the subassemblage may be obtained from Eqs. (14) through
(17) by setting α and β to zero.
The DPF for the joint region of the subassemblage consists
of three parts: shear in the panel, column flexure in the panel,
and beam flexure in the panel. The DPF due to shear is based on
the horizontal shear force in the column part of the joint. Using
Eq. (9) and its virtual counterpart

1
Z β H VC (1−α−β) V̂C (1−α−β)
β β
∇ δJ,S = dx
V̂C 0 GαLt p
V̂C VC H (1 − α − β)2
= . (18a)
Q̂ αβG Lt p
Eq. (18a) may be simplified somewhat by multiplying the
numerator and denominator by H , and by recognizing the
product αLβ H t p as the volume of the panel, v p resulting in
Fig. 8. Effective section for column joint flexure.
V̂C VC H 2 (1 − α − β)2
∇ δJ,S = . (18b) in Fig. 7(a), and the moment of inertia is computed using Eq.
Q̂ Gv p
(19a). When the continuity plates are not present, the effective
Note that there is no additional contribution from the vertical section is as shown in Fig. 7(b), and the moment of inertia
beam shear as expressed in Eq. (10) because the shear stresses is computed from Eq. (19b). If doubler plates are used, they
in the beam–column joint region can only be counted once. should be included in the computation of effective moment of
In fact, Eq. (18) could have been derived using the beam inertia within the joint. This is shown schematically in Fig. 7,
joint shear. This was not done, however, because the joint is and mathematically in Eqs. (19a) and (19b) by the addition of
physically part of the column, not the beam. the term IBDP . It is noted that the schematic representation of
Derivation of the DPF due to joint flexure is based on the the doubler plate in Fig. 7 is somewhat different than may be
moments and shears shown in Figs. 5 and 6 for the column used in practice, where for example, the plate extends above
and beam, respectively. Before the derivations are presented, and below the beam.
it is noted that there is significant uncertainty as to the proper For joint column flexure, the effective cross section is shown
moment of inertia to use in the joint region. For beam joint in Fig. 8, and the related moment of inertia, in absence of dou-
flexure, the moment of inertia depends on whether or not bler plates, is taken to be equal to IC . As with the beam, doubler
beam flange continuity plates are present. When such plates are plates should be included when present. The plate is shown in
present, the effective bending section is assumed to be as shown Fig. 8, and is represented by the term ICDP in Eq. (19c).
F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100 93

Table 2
Ratios of actual to simplified DPFs for column and beam joint flexure

Columns Beams
β = 0.10 β = 0.15 β = 0.20 β = 0.25 β = 0.30 β = 0.10 β = 0.15 β = 0.20 β = 0.25 β = 0.30
α = 0.10 1.12 1.13 1.14 1.15 1.16 1.12 1.19 1.27 1.35 1.44
α = 0.15 1.19 1.21 1.22 1.24 1.26 1.13 1.21 1.29 1.38 1.48
α = 0.20 1.27 1.29 1.31 1.34 1.37 1.14 1.22 1.31 1.41 1.52
α = 0.25 1.35 1.38 1.41 1.44 1.48 1.15 1.24 1.34 1.44 1.56
α = 0.30 1.44 1.48 1.52 1.56 1.61 1.16 1.26 1.37 1.48 1.61

" #
3 b
d B3 twC tcp cp insets of Figs. 5 and 6. If this is done, it can be shown [12] that
IBJ = +4 + tcp bcp (0.5(d B − tcp ))2 the joint flexure DPFs are as follows:
12 12
+ IBDP (19a) V̂C VC H 3 β(1 − β)2
∇ δJ,C F(simple) = (22)
d 3 twC Q̂ 12E ICJ
IBJ = B + IBDP (19b)
12 V̂C VC H 2 Lα(1 − α)2
ICJ = IC + ICDP . (19c) ∇ δJ,B F(simple) = . (23)
Q̂ 12E IBJ
Another factor to consider in determining the effective Before proceeding, it is interesting to compare the joint
moments of inertia in the joint region is the fact that the flexural contributions based on the Actual or the Simplified
portions of the column and beam outside the joint provide moment diagram within the joint. For columns, the ratio of the
flexural restraint for the joint. This is shown schematically by Actual to Simplified DPF is
the “possible zone of influence” in Figs. 7 and 8. To account for
this effect Downs [12] recommended that a multiplier of 1.5 be 3β(1 − α) + (1 − α − β)2
RATIOCJF = (24)
applied to the effective moments of inertia in the joint region. (1 − α)2
Multipliers were not used in the examples shown in this paper.
and the ratio for beams is
However, the use of the multipliers is investigated in part 2 of
this paper [17]. 3α(1 − β) + (1 − α − β)2
RATIOBJF = . (25)
The column joint flexure DPF is based on the moment shown (1 − β)2
in Fig. 5, where the moment is MCC + VCJ x. Substituting from
Table 2 lists the computed ratios for α and β ranging from
Eqs. (9) and (12), the DPF for column joint flexure is
0.10 to 0.30 in increments of 0.05. As may be seen, there is a
βH
1
Z
2 VC V̂C

Hα (1 − α − β)
2 significant difference in the DPFs for flexure based on the actual
∇ δJ,C F = 2 + x dx and simplified moments in the joint region, particularly for the
Q̂ 0 E ICJ 2 β
larger values of α and β. Without a detailed finite element
which when simplified is analysis of the subassemblage, there is no way to know which
approach produces the more accurate result.
V̂C VC H 3 β (1 − α − β)2
 
∇ δJ,C F = α(1 − β) + . (20)
Q̂ 4E ICJ 3 5. Mechanical joint models

The beam joint flexure DPF is based on the moment shown In both the KJ and SJ models the DPFs for the clear span
in Fig. 6, where the moment is MBC + VBJ x. Substituting from regions of the column and beam are the same as that given in
Eqs. (10) and (13), the DPF for beam joint flexure is Eqs. (14)–(17). And, as shown below, the DPF for the panel
Z αL shear is the same as that given in Eq. (18a) or (18b).
H (1 − α − β) 2
 
1 2 VC V̂C Hβ
∇ δJ,B F = 2 + x dx. The KJ model is shown in Fig. 9. As may be seen, the
Q̂ 0 E IBJ 2 αL model consists of four rigid links, connected at the corners by
Upon solving the integral and simplifying, hinges. Mathematical constraints may be used in lieu of the
links, but links are used herein because the links provide a
V̂C VC H 2 Lα (1 − α − β)2 better conceptual model of the joint. Rotational springs at the
 
∇ δJ,B F = β(1 − α) + . (21) upper left and the lower right provide the necessary stiffness.
Q̂ 4E IBJ 3
The upper left spring accounts for panel zone shear stiffness
Given the uncertainty of the moment distribution and and the lower right spring accounts for column flange stiffness.
effective section properties within the joint, it is reasonable to It is very important to note that the column flange stiffness is
adopt a simpler expression for the joint flexure components, not included in the FJ model, and that the beam–column joint
wherein the moment in the joint varies linearly from zero at bending flexibility is not included in the KJ model. Hence, the
the center of the joint to MCJ or MBJ for the column and beam, KJ model is stiffer than the FJ model. A similar problem exists
respectively. These revised moment diagrams are shown in the for the SJ model.
94 F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100

Table 3
Summary of DPFs for isolated subassemblage

DPF Applicable equation Component


component applicable to model
(Y/N)
FJ KJ SJ
VC H (1−β)
Column shear G A SC Y Y Y
VC H 2 (1−α)
Beam shear LG A S B Y Y Y
VC H 3 (1−β)3
Column 12E IC Y Y Y
flexure
VC H 2 L(1−α)3
Fig. 9. Krawinkler Joint model. Beam flexure 12E I B Y Y Y
VC H 2 (1−α−β)2
Joint shear Gv p Y Y Y
VC H 2 Lα (1−α−β)2
Joint beam 4E IBJ [β(1−α)+ 3 ] Y Y N
flexure
(actual)
VC H 3 β (1−α−β)2
Joint column 4E ICJ [α(1 − β) + 3 ] Y Y N
flexure
(actual)
VC H 2 Lα(1−α)2
Joint beam 12E IBJ Y N N
flexure
(simplified)
VC H 3 β(1−β)2
Joint column 12E ICJ Y N N
flexure
Fig. 10. Scissors Joint model. (simplified)

The KJ model, as originally developed by Krawinkler [5], Given expressions are for isolated subassemblage with V̂C / Q̂ = 1.0.
was slightly different than that depicted in Fig. 9, and was
intended to represent inelastic behavior. In the original model, column flange stiffness. As with the KJ model, the column
the shear stiffness was modeled by a shear membrane (instead flange effect will not be included herein.
of a rotational spring), and the column flange spring did not As shown in Charney and Marshall [16], the required
engage (had zero stiffness) until the panel yields in shear. rotational spring stiffness for the SJ model is
Hence, for linear elastic analysis, the column flange spring had KK J Gv p
no effect. Because all of the analysis reported in the current KSJ = = . (30)
(1 − α − β) 2 (1 − α − β)2
paper is for linear elastic systems, the column flange springs
are not considered. In this case, the beam moment in the spring is
According to Charney and Marshall [16] the required M S J = VC H. (31)
rotational spring stiffness of the KJ model is
Using Eq. (28), and recognizing that the virtual moment is
K K J = αLβ H t p G = v p G. (26) the same as above with V̂C substituted for VC , it is seen that
When the column shear is VC , the moment in the spring is
V̂C VC H 2 (1 − α − β)2
M K J = VCJ β H = V H (1 − α − β). (27) ∇ Sδ J = (32)
Q̂ Gv p
The virtual moment is the same, with V̂C being used in lieu which is exactly the same as for the FJ and KJ models.
of VC . The displacement participation for a rotational spring is For the reader’s convenience, the DPF expressions for all
three of the models are summarized in Table 3. Note that V̂C / Q̂
1 M M̂
∇ Kδ J = . (28) was taken as 1.0 for all expressions listed in the table.
Q̂ K K J
6. Three dimensional finite element analysis
Substituting from Eqs. (26) and (27)

V̂C VC H 2 (1 − α − β)2 To establish the relative accuracy of the derivations


∇ Kδ J = (29) presented above, the subassemblages indicated in Table 1 were
Q̂ Gv p
analyzed using the expressions summarized in Table 3, as well
which is exactly the same as that shown in Eq. (18a) for the FJ as through a detailed three dimensional finite element analysis
model. (FEA). In the FEA, the subassemblages were modeled via 8-
The SJ model is shown in Fig. 10, where it may be seen node three dimensional solid elements. The beam and column
that the column and beam are connected through two rotational of each model consisted of an assemblage of rectangular
springs, one representing panel shear, and the other representing flange and web plates, and as such, fillets were not included
F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100 95

subassemblage drift, part (c) gives the shear stresses in the panel
zone computed with Eq. (11), and part (d) provides the results
from the finite element analysis. Part (e) of the table presents
ratios of the simplified model results to the FEA results.
The columns in parts (a), (b), (c) and (e) of the tables
represent different assumptions used in the simplified analysis.
The two columns with the heading “Using Actual Joint
Moments” computes the joint flexure DPFs using Eqs. (20) and
(21). The two columns under the heading “Using Simplified
Joint Moments” are based on the simplified flexural formula
of Eqs. (22) and (23). Analysis was also run with center-to-
center joint dimensions, or out-to-out joint dimensions. Center-
Fig. 11. One-eighth finite element model. to-center analysis uses the distance between the centers of the
flanges to determine the parameters α and β (see Eqs. (4) and
in the finite element analysis. In order to produce a model (5)), while the out-to-out method uses the total depth.
with maximum accuracy (a highest possible resolution mesh), As may be seen from Table 4(a), the various assumptions
only 1/8 of the total subassemblage was modeled. Boundary used in the analysis have a marginal effect on the total
conditions used in the 1/8 model analysis were verified through deflection, which is in the range of 1.752–1.898 cm. From
the analysis of full subassemblages. The detailed FEA was Table 4(b), it may be seen that the combined shear deformations
performed using the program WoodFrameSolver [22] and was in the clear span of the beams and columns accounts for about
verified using SAP2000 [23]. Fig. 11 shows the 1/8 model for 15% of the total drift, and that shear deformation in the panel
the subassemblage with continuity plates. zone alone is responsible for approximately 28% of the total
While the simplified models are able to provide a breakdown drift. Hence, all sources of shear deformation produce 43%
of the contribution of subassemblage drift to the various of the displacement in the subassemblage. Joint beam flexure
components of the subassemblage and to the different and joint column flexure, taken together, are responsible for
sources of deformation, the FEA approach can only provide about 10–11% of the total drift. While smaller than the joint
total subassemblage drift. Hence, the only displacement shear contribution, this could not be considered as a negligible
comparisons that are made are for total drift. Comparisons are source of deformation. The deflection due to all sources of joint
also provided for panel zone shear stress as provided by the deformation is as high as 39.9% of the total subassemblage
simplified formulas and the FEA method. drift.
It is noted that it is possible to provide more detailed When beam flange continuity plates are included, there is
comparisons between the simplified models and the FEA a reduction in the beam joint flexure component only. This is
approach by extending the DPF concept to the FEA results [20]. shown in Table 5, where it is seen that the contribution by beam
In fact, such a presentation is provided in the companion joint flexure has reduced to an average of about 2%. However,
paper [17] that is published in the same journal in which the the total joint flexural component is still not negligible, at about
current paper appears. 8%.
The results from the finite element analysis are provided
7. Presentation of results in Tables 4(d), 5(d) for models A20n and A20c, respectively.
For model A20n, the FEA drift is 1.768 cm, which is
The results of the analysis are presented in two parts. First, within the range of values (1.752–1.898 cm) computed from
detailed results are presented and discussed for Subassemblages the simplified model. The model using the out-to-out joint
A20n and A20c. This is followed by a summary of the results dimensions produced the best correlation with the FEA.
of all subassemblage analysis, as well as correlation with the The shear stress in the panel zone of Model A20n is 153.0
results of the detailed finite element analysis. MPa for the center-to-center model, and 127.4 MPa for the out-
Subassemblages A20n and A20c have a beam span of 6.10 m to-out model. The FEA shear stress 140.3 MPa, which is within
(20 ft) and a column height of 3.81 m (12.5 ft). The column and the range predicted by the simplified model. Similar agreement
beam sections are W360 × 382 (W14 × 257) and W770 × 220 is seen for subassemblage A20c (see Table 5(a) through 5(d)),
(W30 × 148), respectively. The modulus of elasticity of the and for all of the other models analyzed. It is not apparent
steel is taken as 200.0 GPa (29 000 ksi), and Poisson’s ratio that either approach produces better results than the other. It
is 0.3. The column shear used in the analysis was 444.82 kN is noted, however, that the average shear stress from the two
(100 kips). Section properties used in the analysis were based simplified approaches is almost identical to that computed using
on the overall section dimensions, and did not include fillets. FEA.
This was done to provide consistency with the FEA which also While not the main subject of this paper, it is interesting to
did not include fillets. compute the Sensitivity Indices (SI) for the various components
The results for subassemblage A20n are given in Table 4, of the subassemblage. These indices are equal to a DPF divided
which has five parts. Part (a) shows the actual DPF values, by the volume of material in the component that contributed to
part (b) presents the values as a percentage of the total the DPF. Various SIs (multiplied by 100 000) are presented in
96 F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100

Table 4
Results of analysis of subassemblage A20n

(a) Actual DPF values


Deformation source (cm) Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out
Beam flexure 0.495 0.482 0.495 0.482
Beam shear 0.106 0.105 0.106 0.105
Column flexure 0.377 0.366 0.377 0.366
Column shear 0.164 0.162 0.164 0.162
Joint beam flexure 0.093 0.105 0.074 0.083
Joint column flexure 0.100 0.103 0.092 0.094
Panel zone shear 0.563 0.459 0.563 0.459
Total 1.898 1.782 1.872 1.752

(b) Percent of total


Deformation source (%) Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out

Beam flexure 26.1 27.1 26.4 27.5


Beam shear 5.6 5.9 5.7 6.0
Column flexure 19.9 20.5 20.1 20.9
Column shear 8.6 9.1 8.8 9.3
Joint beam flexure 4.9 5.9 4.0 4.7
Joint column flexure 5.3 5.8 4.9 5.4
Panel zone shear 29.7 25.8 30.1 26.2
Total 100.0 100.0 100.0 100.0

(c) Shear stress


Panel zone shear stress (MPa (ksi)) Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out
153.0 (22.2) 127.4 (18.5) 153.0 (22.2) 127.4 (18.5)

(d) Results from finite element analysis


Total displacement (cm) 1.768
Average panel zone shear stress (MPa) 140.3

(e) Simple model to FEA model ratios


Deformation source (%) Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out

Total displacement 1.073 1.008 1.0294 0.991


Panel shear stress 1.095 0.908 1.095 0.908

Table 6 for subassemblage A20c (using actual joint moments joint flexural deformation is not negligible. It is also clear that
and center-to-center dimensions). As may be seen in the table, the addition of the continuity plates had a marginal effect on
the SI for the shear panel alone is 18 approximately times that reducing the subassemblage displacement. This fact is also
for the column or beam. Based on the idea that an optimum evident from Table 7(b), where it is seen that joint flexure
structure has equal SI for all components [19], it is seen that (Beam + Column) was responsible for about 11% of total drift
adding one kilogram of steel to the panel zone (by adding a when continuity plates are not present, and about 8% of total
doubler plate) is more than 18 times as effective in reducing drift when continuity plates are added.
drift than would be adding the same kilogram of steel to the Similar results are seen from Table 8(a) and (b). However,
column. However, the cost of fabricating and attaching that in this case, the portion of subassemblage drift due to joint
extra kilogram of steel needs to be considered as well. It might flexure has increased to as much as 15% for Subassemblage
be more economical to increase the size of the entire column B15n. This drops to about 8% when the continuity plates are
(picking a column with a thicker web), than it would be to add added. Hence, the continuity plates appear to be more effective
a doubler plate. in reducing displacements when the column is deeper. In this
A summary of the results for all “A” subassemblages is case, the reduction in drift associated with the addition of the
presented in Table 7. Similar results are presented in Table 8 continuity plates is 7%.
for the “B” subassemblages. In each case, the analysis used the Tables 7(c) and 8(c) compare the drifts computed using
simplified expressions for moments in the joint region and the the simplified formulas with those found from detailed finite
out-to-out joint dimensions. element analysis. For the A subassemblages, the agreement
As seen in Table 7(a), the component of drift due to beam is excellent, with the maximum difference being only about
column joint deformation is very significant in all cases, and 2.4%. Differences are slightly greater for the B subassemblages,
F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100 97

Table 5
Results of analysis of subassemblage A20c

(a) Actual DPF values


Deformation source (in.) Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out
Beam flexure 0.495 0.482 0.495 0.482
Beam shear 0.106 0.105 0.106 0.105
Column flexure 0.377 0.366 0.377 0.366
Column shear 0.164 0.162 0.164 0.162
Joint beam flexure 0.035 0.039 0.028 0.031
Joint column flexure 0.100 0.103 0.092 0.094
Panel zone shear 0.563 0.459 0.563 0.459
Total 1.839 1.717 1.825 1.700

(b) DPFs as percent of total


Deformation source (%) Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out

Beam flexure 26.9 28.1 27.1 28.4


Beam shear 5.8 6.1 5.8 6.2
Column flexure 20.5 21.3 20.7 21.5
Column shear 8.9 9.4 9.0 9.5
Joint beam flexure 1.9 2.3 1.5 1.8
Joint column flexure 5.4 6.0 5.1 5.5
Panel zone shear 30.6 26.7 30.9 27.0
Total 100.0 100.0 100.0 100.0

(c) Shear stress


Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out

Panel zone shear stress (MPa (ksi)) 153.0 (22.2) 127.4 (18.5) 153.0 (22.2) 127.4 (18.5)

(d) Results from finite element analysis


Total displacement (cm) 1.738
Average shear stress (MPa) 141.0

(e) Simple model to FEA model ratios


Deformation source (%) Using actual joint moments Using simplified joint moments
Center to center Out to out Center to center Out to out

Total displacement 0.945 0.988 1.050 0.978


Panel shear stress 1.085 0.903 1.085 0.903

Table 6 is very significant. Clearly, analytical models that assume


Sensitivity indices for subassemblage A20c that the entire joint region is “rigid” will underestimate the
Item Volume DPF (cm) 100 000∗ S I S IPanel /S I lateral displacements. When centerline analysis is used, the
(cm3 ) flexural deformations in the beam–column region will be
Total 690 452 1.839 0.266 12.86 overestimated, and the shear deformations in the joint region
Beam 318 794 0.601 0.188 18.18 will be underestimated. This may be seen from Fig. 5, where
Column 295 656 0.541 0.183 18.72 the column force diagrams are shown for the fictitious joint
Joint 75 970 0.698 0.918 3.73
model. If centerline analysis were used, the shear, Vc , would be
Panel only 16 452 0.563 3.425 1
constant through the joint, and the bending moment would be
linear from the inflection point of the column to the centerline
where the maximum difference is 3%. Shear stresses computed of the joint.
from the simple and detailed analysis are shown in Tables 7(d)
Considering subassemblage A20n, for example, the center-
and 8(d). Again, the agreement is quite good with differences
line displacements are as follows:
being consistently less than 10%.
Beam Flexure: 0.596 cm
8. Use of centerline analysis in lieu of explicit joint Beam Shear: 0.112 cm
modeling Column Flexure: 0.727 cm
This paper has shown that the influence of deformations Column Shear: 0.204 cm
in the beam–column joint region of moment frame structures Total: 1.639 cm.
98 F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100

Table 7
Summary of A-model DPFs from simplified subassemblage analysis using simplified moments and out-to-out dimensions

(a) Displacement participation factors


Deformation source Model
A15n A20n A25n A15c A20c A25c
Beam flexure 0.336 0.482 0.630 0.336 0.482 0.630
Beam shear 0.137 0.105 0.085 0.137 0.105 0.085
Column flexure 0.366 0.366 0.366 0.366 0.366 0.366
Column shear 0.162 0.162 0.162 0.162 0.162 0.162
Joint beam flexure 0.079 0.083 0.085 0.029 0.031 0.032
Joint column flexure 0.094 0.094 0.094 0.094 0.094 0.094
Panel zone shear 0.431 0.459 0.477 0.431 0.459 0.477
Total 1.604 1.752 1.899 1.555 1.700 1.846

(b) DPFs as percent of total


Deformation source Model
A15n A20n A25n A15c A20c A25c

Beam flexure 20.9 27.5 33.2 21.6 28.4 34.1


Beam shear 8.5 6.0 4.5 8.8 6.2 4.6
Column flexure 22.8 20.9 19.3 23.5 21.5 19.8
Column shear 10.1 9.3 8.5 10.4 9.5 8.8
Joint beam flexure 4.9 4.7 4.5 1.9 1.8 1.7
Joint column flexure 5.9 5.4 5.0 6.1 5.5 5.1
Panel zone shear 26.9 26.2 25.1 27.7 27.0 25.8
Total 100.0 100.0 100.0 100.0 100.0 100.0

(c) Total displacements


Item Model
A15n A20n A25n A15c A20c A25c

Simple displacement 1.604 1.752 1.899 1.555 1.700 1.846


FEA displacement 1.623 1.768 1.915 1.594 1.738 1.883
Simple/FEA 0.988 0.991 0.992 0.976 0.978 0.980

(d) Shear stresses in panel zone


Item Model
A15n A20n A25n A15c A20c A25c

Simple shear stress 123.4 127.4 129.8 123.4 127.4 129.8


FEA shear stress 136.3 140.3 142.7 137.0 141.0 143.4
Simple/FEA 0.905 0.908 0.910 0.901 0.904 0.905
All DPFs in cm units, all displacement is cm units, and all shear stresses in MPa units.

This total is less than that produced by any of the mechanical 9. Summary, conclusions, and recommendations
models (see Table 4(a)), indicating that even centerline analysis
is unconservative (too stiff) with respect to predicting drift. In this paper, three simplified models for computing
It is important to note, however, that the rigid end zone displacement participation factors (DPFs) in beam–column
analysis of the same subassemblage would produce a total joint subassemblages were derived. The models presented were
the Fictitious Joint (FJ) model, the Krawinkler Joint (KJ)
drift of 1.142 cm, which is only 63% of the displacement
model, and the Scissors Joint (SJ) model. The FJ model
produced by the models that explicitly include beam–column
includes the effect of flexural deformations in the beam–column
joint deformation.
joint, but the KJ and SJ models do not.
While centerline analysis is somewhat unconservative in the Based on the results obtained from the analysis presented
example given, it is clearly an improvement over rigid end zone herein, and recognizing that only twelve individual subassem-
analysis. Hence, centerline analysis should be used whenever blages were analyzed, the following conclusions and recom-
the more accurate modeling approaches are not available. mendations are made.
For existing buildings that were analyzed using rigid end 1. When compared to the results of detailed finite element
zones, it is clear that the lateral drift will be underestimated. analysis, the use of the formulas provided in Table 3 provide
This will have an effect on the serviceability of the structure, reasonably accurate estimates of the displacements in
and may have an adverse effect on the strength as well. Strength beam–column subassemblages. The best agreement appears
would be influenced most significantly for structures which are to be obtained when the out-to-out joint dimensions are used.
sensitive to P-delta effects. The seriousness of such problems Use of actual versus simplified moments in the joint made
can only be assessed on a case-by-case basis. little difference with regards to predicted deflection.
F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100 99

Table 8
Summary of B-model DPFs from simplified subassemblage analysis using simplified moments and out-to-out dimensions

(a) Displacement participation factors


Deformation source Model
B15n B20n B25n B15c B20c B25c
Beam flexure 0.278 0.421 0.566 0.278 0.421 0.566
Beam shear 0.128 0.100 0.082 0.128 0.100 0.082
Column flexure 0.147 0.147 0.147 0.147 0.147 0.147
Column shear 0.108 0.108 0.108 0.108 0.108 0.108
Joint beam flexure 0.126 0.137 0.144 0.043 0.047 0.049
Joint column flexure 0.038 0.038 0.038 0.038 0.038 0.038
Panel zone shear 0.257 0.287 0.305 0.257 0.287 0.305
Total 1.081 1.237 1.389 0.998 1.146 1.294

(b) DPFs as percent of total


Deformation source Model
B15n B20n B25n B15c B20c B25c

Beam flexure 25.7 34.0 40.7 27.9 36.7 43.7


Beam shear 11.9 8.1 5.9 12.8 8.7 6.4
Column flexure 13.6 11.9 10.6 14.7 12.8 11.3
Column shear 9.9 8.7 7.7 10.8 9.4 8.3
Joint beam flexure 11.7 11.1 10.4 4.3 4.1 3.8
Joint column flexure 3.5 3.0 2.7 3.8 3.3 2.9
Panel zone shear 23.8 23.2 22.0 25.7 25.0 23.6
Total 100.0 100.0 100.0 100.0 100.0 100.0

(c) Total displacements


Item Model
B15n B20n B25n B15c B20c B25c

Simple displacement 1.081 1.237 1.389 0.998 1.146 1.294


FEA displacement 1.050 1.202 1.352 1.006 1.153 1.301
Simple/FEA 1.030 1.029 1.028 0.992 0.994 0.995

(d) Panel zone shear stresses


Item Model
B15n B20n B25n B15c B20c B25c

Simple shear stress 79.9 84.4 87.1 79.9 84.4 87.1


FEA shear stress 85.8 90.3 93.0 87.0 91.4 94.3
Simple/FEA 0.931 0.935 0.937 0.918 0.923 0.924
All DPFs in cm units, all displacement is cm units, and all shear stresses in MPa units.

2. The use of center-to-center joint dimensions overestimates The basic conclusion of this paper is that beam–column joint
the average joint shear stress, and the use of centerline flexural deformations are significant, and should not be ignored.
dimensions underestimates the average joint shear stress. As mentioned in the beginning of this paper, such deformations
However, errors in average shear stress were generally less are ignored in both the Krawinkler and Scissors models. In fact,
than 10%. these models typically include an added stiffness component
3. Shear deformations in the beam, the column, and (due to column flange bending) that was not included in the
particularly in the panel zone of the beam–column joint are current analysis.
very significant, and should not be ignored. Because the Krawinkler and Scissors models are most often
4. Deflections due to flexural deformations in the beam–column used in practice, it is desirable to modify these models to
joint are smaller than those due to shear deformations, but account for flexural deformations in the joint. It is suggested in
are not negligible, and should not be ignored. Such defor- Part 2 of this paper that this may be accomplished by providing
mations were responsible for as much as 15% of the drift in some flexural flexibility in the rigid links used in these models.
the subassemblages without beam-flange continuity plates. It is also noted that while the detailed FEA produced total
Addition of such plates reduced the drift by as much as 7% displacements that are very consistent with those obtained
in some subassemblages, and as little as 3% in others. from the simplified methods, there was no way to determine,
5. There is considerable uncertainty regarding the appropriate on a one-to-one basis, how the individual components of
section property to use for computing flexural deformations displacement compare. For example, it may be the case that
in the beam–column joint. This is particularly true for beam the subassemblage models underestimate panel zone shear
flexure when beam-flange continuity plates are not provided. deformation while overestimating joint flexural deformation. In
100 F.A. Charney, R. Pathak / Journal of Constructional Steel Research 64 (2008) 87–100

order to clarify this point, Part 2 of this paper extends the DPF [11] Schneider SP, Amidi A. Seismic behavior of steel frames with deformable
approach to the detailed finite element analysis. This analysis panel zones. Journal of Structural Engineering 1998;124(1):35–42.
[12] Downs WM. Modeling and behavior of beam/column joint regions of
shows that the simplified model consistently overestimates the
steel frames. M.S. thesis. Blacksburg (VA): Department of Civil and
displacement due to flexure in the joint, although the errors Environmental Engineering, Virginia Tech; 2002.
produced are generally small. [13] Foutch DA, Yun S. Modeling of steel moment resisting frames for seismic
loads. Journal of Constructional Steel Research 2002;58:529–64.
References [14] Kim K, Engelhardt MD. Monotonic and cyclic loading models for panel
zones in steel moment frames. Journal of Constructional Steel Research
[1] Bertero VV, Popov EP, Krawinkler H. Beam–column subassemblies 2002;56:605–35.
under repeated load. Journal of the Structural Division 1972;98(ST5): [15] Gupta A, Krawinkler H. Relating seismic drift demands of SMRFs to
1137–59. element deformation demands. Engineering Journal 2002;39(2):100–8.
[2] Becker R. Panel zone effect on the strength and stiffness of steel rigid [16] Charney FA, Marshall JA. Comparison of the Krawinkler and Scissors
frames. Engineering Journal 1975;19–29. models for including beam–column joint deformations in the analysis of
[3] Krawinkler H, Bertero VV, Popov EV. Shear behavior of steel frame moment resisting frames. Engineering Journal 2006;43(1):31–48.
joints. Journal of the Structural Division 1975;101(11):2317–36. [17] Charney FA, Pathak R. Sources of elastic deformations in steel frame and
[4] AISC. Tests and analysis of panel zone behavior in beam-to-column framed-tube structures: Part 2: Detailed subassemblage models. Journal
connections. American Institute of Steel Construction and the Structural of Constructional Steel Research 2008;64(1):101–17.
Steel Educational Council; 1980. [18] Charney FA. The use of displacement participation factors in the
[5] Krawinkler H. Shear in beam–column joints in seismic design of frames. optimization of wind drift controlled buildings. In: Proceedings of the
Engineering Journal 1978;15(3):82–91. second conference on tall buildings in seismic regions. Council of Tall
[6] Charney FA, Johnson R. The effect of panel zone deformations on the drift Buildings and Urban Habitat; 1991. p. 91–8.
of steel framed structures. In: Presented at the ASCE structure congress. [19] Charney FA. Economy of steel frames through identification of structural
1986. behavior. In: Proceedings of the national steel construction conference.
[7] Krawinkler H, Mohassebs. Effects of panel zone deformation on seismic 1993. p. 12-1–13-33.
response. Journal of Constructional Steel Research 1987;8:233–50. [20] Cowper GR. The shear coefficient in Timoshenko’s beam theory. Journal
[8] Charney FA. Sources of elastic deformations in laterally loaded steel of Applied Mechanics 1966;33:335–40.
frame and tube structures. In: Proceedings of the fourth world congress [21] Charney FA, Iyer H, Spears PW. Computation of major axis shear
on tall buildings. 1990. p. 893–915. deformations in wide flange steel girders and columns. Journal of
[9] Horvilleur JF, Charney FA. The effect of beam–column joint deformation Constructional Steel Research 2005;61(11):1525–58.
on the lateral response of steel frame building structures. In: III Simposio [22] Pathak R. Development of finite element analysis modeling mesh
Internacional Y VI Simposio Natcional de Estructueas de Acero. 1993. p. generation and analysis software for light wood frame houses. Master of
269–303. Science thesis. Blacksburg (VA): Department of Civil and Environmental
[10] Kim K, Englehardt MD. Development of analytical models for earthquake Engineering, Virginia Tech; 2004.
analysis of steel frames. Austin (TX): PMFSEL 95-2, Phil M. Ferguson [23] SAP2000. Integrated software for structural analysis and design, v7. 44.
Structural Engineering Lab, University of Texas; 1995. Berkeley: Computers and Structures, Inc.; 2000.

Anda mungkin juga menyukai