You are on page 1of 31

Journal of Constructional Steel Research 58 (2002) 605–635

www.elsevier.com/locate/jcsr

Monotonic and cyclic loading models for panel


zones in steel moment frames
Kee Dong Kim a, Michael D. Engelhardt b,*
a
Department of Civil Engineering, Kongju National University, Kongju, Chungnam, South Korea
b
Department of Civil Engineering, University of Texas at Austin, Austin, TX 78712-1076, USA

Received 14 June 2001; received in revised form 17 August 2001; accepted 5 October 2001

Abstract

This paper presents the development of analytical models to predict the elastic and inelastic
response of the panel zone portion of columns in steel moment resisting frames. In many
practical cases, the panel zone can dominate the inelastic response of a moment frame, and
accurate panel zone models are needed to realistically predict overall frame performance. Simi-
lar to previous models, the newly proposed models are based on the concept of representing
the panel zone as a nonlinear rotational spring. These new models build on previously
developed models, and introduce a number of features and refinements that show better corre-
lation with available experimental data. The model for monotonic loading is based on quadri-
linear panel zone moment–deformation relations. In this model, both bending and shear defor-
mation modes are considered. The model proposed for cyclic loading is based on Dafalias’
bounding surface theory combined with Cofie’s rules for movement of the bound line.
Additional modifications are suggested to the cyclic loading model to account for the influence
of a composite floor slab on panel zone response. Extensive comparisons with experimental
data are presented.  2002 Elsevier Science Ltd. All rights reserved.

Keywords: Seismic; Inelastic; Bounding surface model; Composite; Connections; Joints

* Corresponding author. Tel.: +1-512-471-6837; fax: +1-512-471-1944.


E-mail address: mde@mail.utexas.edu (M.D. Engelhardt).

0143-974X/02/$ - see front matter  2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 3 - 9 7 4 X ( 0 1 ) 0 0 0 7 9 - 7
606 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

1. Introduction

Mathematical models are developed in this paper for describing the monotonic
and cyclic load–deformation response of the panel zone region of beam–column
joints of a steel moment resisting frame. The panel zone is the portion of the column
contained within the beam–column joint. When a moment frame is subject to lateral
loads, high shear forces develop within the panel zone. The resulting deformations
of the panel zone can have an important effect on the response of the frame in both
the elastic and inelastic ranges of frame behavior [1,2].
Numerous tests have been performed in the past three decades to investigate the
load–deformation behavior of the joint panel using connection subassemblies [3–8].
Some significant observations from these tests are:

앫 Joint panel zones often develop a maximum strength that is significantly greater
than the strength at first yield. This additional strength has been attributed to strain
hardening and to contributions of the column flanges in resisting panel zone shear
forces. Large inelastic panel zone deformations are typically required in order to
develop the maximum panel zone strength.
앫 Panel zone deformations can add significantly to the overall deformation of a steel
moment frame, for both elastic and inelastic ranges of behavior.
앫 Panel zone stiffness and strength can be increased by the attachment of web
doubler plates to the column within the joint region. The effectiveness of doubler
plates is affected by the method used to connect them to the column.
앫 In the inelastic range, panel zones can exhibit very ductile behavior, both for
monotonic and cyclic loading. Experimentally observed hysteresis loops are typi-
cally very stable, even at large inelastic deformations.
앫 Large inelastic panel zone deformations can increase the likelihood of fracture
occurring in the region of the beam flange to column flange groove welds. This
effect has been attributed to the occurrence of large localized deformations or
‘kinks’ in the column flanges at the boundaries of the panel zone.

Current US building code provisions [9–11] permit the formation of plastic hinges
in the panel zones of steel moment frames under earthquake loading. Thus, rather
than forming plastic flexural hinges only in the beams or columns, a primary source
of energy dissipation in a steel moment frame can be the formation of plastic shear
hinges in the panel zones. Consequently, to accurately predict the response of a steel
moment frame under earthquake loading, an accurate analytical model is needed to
predict the response of the panel zone.
To include panel zone deformation in frame analysis, the traditional center-to-
center line representation of the frame must be modified. Fig. 1 shows a comparison
between an experiment on a beam–column subassemblage and an analytical predic-
tion of the subassemblage response. The experiment was specimen A1 reported by
Krawinkler et al. [4]. Inelastic response of this specimen was dominated by yielding
of the panel zone. The analytical results are obtained from a model of the specimen
using center-to-center line dimensions. A beam–column element with plastic hinges
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 607

Fig. 1. Comparison of test results [4] and analytical results obtained by using center-to-center line dimen-
sion modeling.

was employed in this analysis. From the figure it can be seen that the analysis using
the center-to-center line dimension, without explicit modeling of the panel zone, may
produce misleading results. Clearly, the panel zone response must be explicitly mod-
eled to obtain realistic predictions of the overall frame behavior.
To model the behavior of panel zones in frame analysis, Lui [12] developed a
joint model based on the finite element method. The model consists of seven elements
for interior beam-to-column joints: one web element, two flange elements (beam
elements), and four connection elements. Although capable of representing a variety
of deformation modes of panel zones, this model employed a simple hardening rule
suitable for monotonic loading and does not realistically model cyclic behavior.
Another disadvantage of this model is its high computational cost. Other finite
element models using more sophisticated hardening rules could be developed for the
analysis of column panel zones. However, in this study, nonlinear rotational springs
are used as the basis for modeling the panel zone for nonlinear dynamic analysis of
moment resisting frames because of simplicity and computational efficiency.
Several researchers, including Fielding and Huang [13], Krawinkler et al. [4] and
Wang [14] proposed relationships between panel zone shear force V and panel zone
deformation g for monotonic loading. These relationships have been used as the basis
of mathematical models for nonlinear rotational springs representing the panel zone.
Krawinkler’s V–g relations have been adopted in several building codes [9,10] as a
basis for computing the shear strength of panel zones. However, it was pointed out
by Krawinkler that a new model might be needed for joints with thick column flanges
since his V–g relations were derived from experimental and analytical results for
panel zones with relatively thin column flanges. Wang also showed that Krawinkler’s
V–g relations may overestimate panel zone shear strength for panel zones with thick
column flanges.
608 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

In the remainder of this paper, existing analytical models for panel zones will be
examined by making comparisons with experimental data or with other analytical
results. A number of refinements and improvements to existing models are then
suggested to overcome some of the shortcomings of existing models and to provide
better correlation with available experimental data.

2. General characteristics of panel zone element

The panel zone element is essentially a rotational spring element, which transfers
moment between the columns and beams framing into a joint [15]. The panel element
has no dimensions and connects two nodes with the same coordinates. One of these
nodes is attached to the elements modeling the columns framing into the joint, as
shown in Fig. 2, while the other node is attached to the elements modeling the beams.
Therefore, the moment transferred by the panel element is related to the relative
rotation between the columns and beams framing into a joint. The vertical and hori-
zontal translations of the two nodes are constrained to be identical. Therefore, one
vertical, one horizontal, and two rotational degrees of freedom exist at each joint.
The relative rotation between the connected nodes is related to the node rotations
as follows:

dg ⫽ {1 ⫺1} 再 冎
dqI
dqJ
(1)

or
dg ⫽ a·dr (2)
where dg is the increment of relative rotation, which is the panel element defor-
mation, and dqI, and dqJ, are the increments of rotation of the connected nodes.
Then the tangent stiffness relationship for the panel zone element is

Fig. 2. Idealization of beam-to-column joint.


K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 609

dMpa ⫽ Ktdg (3)


pa
where dM is the increment of moment applied on a joint and Kt is the rotational
tangent stiffness of the joint. In terms of nodal rotations, the stiffness, KT, is given by
KT ⫽ aTKta (4)
The definable properties of the panel zone element are the rotational stiffnesses
and yield moments for monotonic loading, and hysteretic rules for cyclic loading.
In the following sections, existing models for monotonic loading are examined and
a new model is presented. This is followed by the development of hysteretic rules
for a cyclic loading model.

3. Review of existing models for monotonic loading

Existing mathematical models for panel zone response under monotonic loading
are typically based on a computation of an approximate equivalent shear force acting
on the panel zone. The boundary forces on a joint panel, shown in Fig. 3, can be
transformed into an approximate equivalent shear force from equilibrium as follows:
Mbl ⫹ Mbr (Vct ⫹ Vcb) Mbl ⫹ Mbr (1⫺r) pa
Veq ⫽ ⫺ ⬇ (1⫺r) ⫽ M (5)
db⫺tbf 2 db⫺tbr db⫺tbf
where r ⫽ (db⫺tbf) / Hc, Mpa ⫽ Mbl ⫹ Mbr is the panel zone moment, tbf is the thick-
ness of the beam flange, db is the beam depth, and Hc is the column height. A key
simplification in this analysis is that the beam moments are replaced by an equivalent
couple, with the forces acting at mid-depth of the beam flanges. These forces produce
a large shear in the panel zone. The shear in the column segments outside of the
panel zone are then subtracted to obtain the net shear force, Veq, acting on the panel
zone. In obtaining the shear forces in the column segments outside of the panel zone,
it is often assumed that points of inflection in the column occur at a distance Hc/2
above and below the panel zone.

Fig. 3. Boundary forces and equivalent shear forces on panel zone.


610 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Panel zone shear force V versus panel zone deformation g relations for monotonic
loading, which are based on the equivalent shear force, Veq, can be transformed into
panel zone moment Mpa versus panel zone deformation g relations by Eq. (5). Field-
ing and Huang [13] proposed a bilinear relationship consisting of an elastic stiffness
Ke followed by a post elastic stiffness K1. Krawinkler et al. [4] and Wang [14]
each proposed different tri-linear Mpa–g relations consisting of an elastic stiffness Ke
followed by two linear post elastic stiffness values K1 and K2. For all three models,
the post elastic stiffness K1 is related to the contribution of the column flanges to
the panel stiffness. For the tri-linear models, the second post elastic stiffness K2 is
associated with strain hardening.
Past researchers computed the elastic stiffness of the panel element by considering
pure elastic shear deformation of an effective shear area of the panel zone. Fielding
and Krawinkler considered the effective shear area Aeff equal to (dc⫺tcf)tcw, and Wang
considered the effective shear area Aeff of (dc⫺2tcf)tcw, where the subscripts ‘c’, ‘f’,
and ‘w’ stand for column, flange, and web, respectively. They suggested the yield
moment and elastic stiffness of the panel zone be taken as follows:
Vydb t̄yAeffdb
y ⫽
Mpa ⫽ (6a)
(1⫺r) (1⫺r)
Mpa GAeffdb
Ke ⫽ ⫽
y
(6b)
gy (1⫺r)
where Vy is the yield shear force of the panel zone, gy ⫽ t̄y / G, G is the elastic shear
modulus, and t̄y is the Von Mises yield shear stress of the column web, based on
shear and axial force interaction. The Von Mises yield shear stress, t̄y, is taken as:

t̄y ⫽
sy
冑1⫺(P / P ) 2

冑3 y (7)

where P and Py are the axial force and the axial yield force on the column, respect-
ively, and sy is the yield stress of the column web.
For the inelastic range, Fielding considered a bilinear model with the following
post-elastic stiffness K1:
5.2Gbcft3cf
K1 ⫽ (8)
db(1⫺r)
where bcf and tcf are width and thickness of column flange, respectively. Krawinkler
proposed empirical formulas for the post-elastic stiffness K1 and the second yield
moment Mpa sh as follows:

1.04Gbcft2cf
K1 ⫽ (9)
(1⫺r)
3.12t̄ybcft2cf
sh ⫽ My ⫹
Mpa pa
(10)
(1⫺r)
Wang suggested the post-elastic stiffness K1, as follows:
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 611

K1 ⫽ 0.7Gbcft2cf (11)
Krawinkler and Wang assumed that strain hardening begins at gsh ⫽ 4gy and 3.5gy,
respectively. The strain-hardening branch stiffness K2 was suggested as follows:
GstAeffdb
K2 ⫽ (12)
(1⫺r)
where Gst is the strain hardening shear modulus.
The existing models described above are compared to four specimens tested by
Krawinkler et al. [4], Fielding and Huang [13], and Slutter [5] in Figs. 4–7. Each
of these specimens was a beam–column subassemblage with a weak panel zone. Fig.
4 shows Krawinkler et al.’s specimen A2, which had a column flange thickness of
1 cm. Slutter’s specimen 1, which had a column flange thickness of 1.8 cm, is plotted
in Fig. 5. Krawinkler et al.’s specimen B2, with a column flange thickness of 2.37
cm, is shown in Fig. 6. Finally, Fig. 7 shows Fielding and Huang’s test specimen,
which had a column flange thickness of 3.5 cm.
Additional comparisons are shown in Figs. 8–12. In these figures, FEM analysis
predictions are provided for Slutter’s specimen 1 and compared to the simplified
nonlinear spring models. This specimen was analyzed a number of times, varying the
column flange thickness. These FEM analyses provide an indication of the expected
response of panel zones as the column flange thickness varies, but where all other
variables remain constant. These finite element analyses were reported by Wang
[14]. The specimen tested by Slutter was analyzed by using a two-dimensional FEM,
in which the flanges and webs were represented by beam elements and plain stress
elements, respectively. The finite element results are compared with the correspond-
ing test data in Fig. 5, for the actual specimen with a 1.8 cm thick column flange.

Fig. 4. Comparison of the monotonic model and test data for Krawinkler et al.’s [4] specimen A2 with
tcf ⫽ 1 cm.
612 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 5. Comparison of the monotonic model and test data for Slutter’s [5] specimen 1 with tcf ⫽
1.8 cm.

Fig. 6. Comparison of the monotonic model and test data for Krawinkler et al.’s [4] specimen B2 with
tcf ⫽ 2.37 cm.

This comparison shows that Wang’s finite element analysis reasonably predicted the
observed panel zone behavior. Figs. 8–12 show finite element predictions for a model
based on Slutter’s specimen, but considering column flange thickness values equal
to 0.9, 1.35, 2.7, 3.6 and 4.51 cm. In each figure, the corresponding predictions of
the simple nonlinear spring elements are also plotted.
A number of observations can be made from the comparisons plotted in Figs. 4–
12. Fielding’s bilinear model performs well for small panel zone rotations, but this
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 613

Fig. 7. Comparison of the monotonic model and test data for Fielding and Huang’s [13] specimen with
tcf ⫽ 3.5 cm.

Fig. 8. Comparison of the monotonic model and FEM results for Slutter’s specimen 1 with tcf ⫽
0.9 cm.

model shows rather poor performance at large rotations regardless of column flange
thickness because the model neglects strain-hardening effects. The performance of
Krawinkler et al.’s model appears reasonable for panel zone joints with column
flange thickness less than about 2.5 cm. However, for thicker column flanges, this
model somewhat overestimates panel zone strength. Wang’s model generally under-
estimates panel zone strength regardless of column flange thickness, apparently
because in this model the effective shear area of the panel zones is calculated as
(dc⫺2tcf)tcw instead of the other models’ effective shear area (dc⫺tcf)tcw.
614 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 9. Comparison of the monotonic model and FEM results for Slutter’s specimen 1 with tcf ⫽
1.35 cm.

Fig. 10. Comparison of the monotonic model and FEM results for Slutter’s specimen 1 with tcf ⫽
2.7 cm.

4. Proposed model for monotonic loading and comparison with test data

As the ratio of column flange thickness to column depth increases, the influence
of column flange thickness on panel zone yield moment and elastic stiffness increases
[14]. Panel zone models that include shear deformations only cannot account for this
increase in yield moment and elastic stiffness according to the increase in the ratio
of column flange thickness to column depth. Thus, in this study, both bending and
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 615

Fig. 11. Comparison of the monotonic model and FEM results for Slutter’s specimen 1 with tcf ⫽
3.6 cm.

Fig. 12. Comparison of the monotonic model and FEM results for Slutter’s specimen 1 with tcf ⫽
4.51 cm.

shear deformation modes are included in the panel zone model. The resulting mono-
tonic model has quadri-linear Mpa–g relations.
It is assumed that the panel zone can be considered as two equivalent beams,
which are symmetric with respect to the center of the panel zone and are fixed at
the center. The boundary condition at the other end of these beams is considered to
be somewhere between free and fixed. Thus, the displacement of the equivalent beam
due to the shear force Veq can be described as follows:
616 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

⌬⫽ 冉 1

1
V冊
S1 S2 eq
(13)

where 1/S1 is the bending flexibility of the equivalent beam and 1/S2 is the shear
flexibility. The bending and shear flexibilities can be described as follows:
1 [(db⫺tbf) / 2]3 1 [(db⫺tbf) / 2]
⫽ and ⫽ (14)
S1 CrEI S2 G(dctcw ⫹ RfAdp)
where I is the moment of inertia of the column section, Cr is a constant to be determ-
ined according to the degree of end restraint, Adp is the area of a doubler plate, and
Rf is the reduction factor to account for the strain incompatibility between a doubler
plate and column web. Eq. (13) can be rewritten in terms of the panel zone moment
Mpa and the rotation g as follows:
S1S2 S1S2 (db⫺tbf)
Veq ⫽ ⌬⫽ g (15a)
S1 ⫹ S2 S1 ⫹ S2 2
(db⫺tbf) S1S2 (db⫺tbf)(db⫺tbf)
Mpa ⫽ Veq ⫽ g (15b)
(1⫺r) S1 ⫹ S2 2 (1⫺r)
Thus, the elastic stiffness is
S1S2 (db⫺tbf)(db⫺tbf)
Ke ⫽ (16)
S1 ⫹ S2 2 (1⫺r)
The yield moment of panel zone is defined as follows:
y ⫽ KeCygy
Mpa (17)
where Cygy is the average shear deformation of panel zone at which shear yielding
occurs and Cy is the ratio of the average shear deformation to gy.
To describe the behavior of the panel zone in the range from first shear yielding
up to the entire shear yielding of the panel zone, it is assumed that for this range
the panel zone can be modeled as two separate beams with T-shaped sections, similar
to Fielding and Huang’s approach [13]. The web depth of the T-shaped section is a
quarter of the column web depth. Then, the post-elastic stiffness K1 can be defined as
PS1PS2 (db⫺tbf)(db⫺tbf)
K1 ⫽ 2 (18a)
PS1 ⫹ PS2 2 (1⫺r)
CrEIT G[(dc / 2⫺dyw)tcw ⫹ RfAdp / 4]
PS1 ⫽ and PS2 ⫽ (18b)
[(db⫺tbf) / 2]3 [(db⫺tbf) / 2]
where IT is the moment of inertia of the T section and dyw is a quarter of the web
depth. The second yield moment Mpa y1 is

(db⫺tbf)
y1 ⫽ t̄y(dctcw ⫹ RfAdp)
Mpa (19)
(1⫺r)
The second post-elastic stiffness after shear yielding of the entire panel zone is
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 617

defined by using an approach similar to Krawinkler et al. [4]. The panel zone, after
the shear yielding of the entire web, consists of an elastic–perfectly plastic shear
panel surrounded by rigid boundaries with springs at the four corners. It is assumed
that these springs simulate the resistance of elements surrounding the panel, in parti-
cular the bending resistance of column flanges, and that the spring stiffness can be
approximated by
Ebcft2cf
Ks ⫽ (20a)
Cs
where Cs is a constant to be determined from test results. From the work equation
and Eq. (20a), the second post-elastic stiffness K2 is obtained as
4Ebcft2cf
K2 ⫽ (20b)
Cs(1⫺r)
It is assumed that when strain hardening starts, plastic hinges form in the column
flanges at the four corners of the panel [14]. Then, the third yield moment of the
panel zone at which strain hardening initiates can be defined as
y2 ⫽ My1 ⫹ s̄ybcftcf
Mpa pa 2
(21a)
s̄y ⫽ s (1⫺(P / Py) )
fl
y
2
(21b)
where sfly is the yield stress of the column flange.
The strain hardening stiffness K3 is
Gst(dctcw ⫹ RfAdp)(db⫺tbf)
K3 ⫽ (22)
(1⫺r)
This new proposed monotonic model has been applied to the same four specimens
as in the previous section. Figs. 4–12 show comparisons of this new model with test
results and FEM results for various values of column flange thickness. Based on
calibration to test results (Figs. 4–7), it has been found that Cr ⫽ 5, Cy ⫽ 0.8–0.9,
and Cs ⫽ 20 are reasonable. For the comparison of the model and FEM results (Figs.
8–12), the value of Cy is chosen to be Cy ⫽ 1. From the comparison with test results,
it can be seen that the performance of the new model shows a smoother transition
from elastic to inelastic behavior than the other models because it has quadri-linear
relations. From the comparison with FEM results, it can be seen that the new model
can reasonably describe the increase of yield strength and elastic stiffness according
to an increase in the ratio of column flange thickness to column depth due to the
inclusion of the bending deformation mode. The suggested model significantly under-
estimates the panel zone strength of Krawinkler et al.’s specimen B2 (Fig. 6). This
specimen exhibited unusually early strain hardening due to a very short yield plateau
(es ⫽ 4.4ey) of the stress–strain relation of the column web material [4]. In fact, all
of the models shown in Fig. 6 underestimated the strength of this specimen. However,
for the remainder of the comparisons, the correlation between the predictions of this
new model and the response obtained by test or FEM analysis is quite good regard-
less of column flange thickness.
618 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

In this study, the thickest column flange used to evaluate panel zone models was
tcf ⫽ 4.5 cm. In actual design practice, even thicker column flanges may be used,
perhaps on the order of 8–13 cm. Additional test predictions for such column sections
are needed to further verify this monotonic model and to investigate maximum shear
strength. No such data were found in the literature.

5. Proposed model for cyclic loading

As noted earlier, panel zone response can have a particularly important effect on
the behavior of steel moment frames under earthquake loading. Consequently, accur-
ate hysteretic rules are needed to predict panel zone response under cyclic and ran-
dom loading. A common model used in the past is based on bilinear kinematic
hardening, and this model has been widely used for inelastic dynamic analysis of
moment resisting frames. Fig. 21 shows a comparison of the bilinear kinematic hard-
ening model and test data for the panel zone of Krawinkler et al.’s specimen A1
[4]. A similar comparison is shown in Fig. 27 for the overall subassemblage response
for this test specimen. From both figures, it is clear that the response predictions of
the bilinear model do not correlate well with the experimentally observed response.
The bilinear model substantially underestimated panel zone strength in the latter
cycles of loading. The correlation in the later loading cycles could be improved by
increasing the yield strength of the bilinear model, but then the correlation would
be particularly poor for the early loading cycles. In general, the ability of the bilinear
model to replicate the full loading history of the panel zone is limited. Consequently,
a new model was developed to provide improved panel zone response predictions
over a wide range of loading histories.
In this study, hysteretic rules for the panel zone are developed based on Dafalias’
bounding surface theory [16]. This model also uses Cofie and Krawinkler’s rules for
the movement of the bound line [17]. Based on observations from experiments and
FEM analyses for panel zones, it has been found that for large plastic rotations, the
shear strains in the panel zone are distributed nearly uniformly within the panel, and
the value of joint rotation is close to the value of the average shear strain in the
panel [4,14]. Therefore, it is assumed that the panel zone moment–rotation relation-
ships can be determined from the material properties of the panel zone using Cofie’s
rules. These rules for the movement of the bound line, which were developed for
stress–strain relationships, will be adopted for the panel zone moment–rotation
relationships.
The main feature of Cofie’s model is that the cyclic steady state curve is used to
describe the movement of the bounding line. In this study, the same kind of cyclic
steady state curve is developed to describe the movement of the bounding line for
the cyclic behavior of the panel zone, as follows:
g Mpa
⫽ pa ⫹
gn Mn
Mpa
xMpa
n
冉 冊 c
(23)

where Mpa
n and gn are the normalizing panel moment and corresponding elastic
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 619

rotation, respectively. By comparison with available cyclic test data, it has been
empirically found that the constant C of the cyclic steady state curve is 7 and x is
1.1–1.2 [2]. The cyclic steady state curve and bound lines are shown together with
experimental data [7] in Fig. 13.
Experimental and FEM results suggest that column flanges do not significantly
influence panel zone stiffness during cyclic loading, but do have a significant effect
on panel zone strength. From FEM results for joints with the same dimensions except
for the column flange thickness, it has been found that the effect of column flange
thickness on the strength of the joint during cyclic loading can be normalized by
Mpa
n [14] as follows

n ⫽ My ⫹ 2Mpcf
Mpa pa
(24)
where Mpcf is the plastic moment of the column flange. The elastic rotation corre-
sponding to the normalizing moment Mpan is

Mpa
gn ⫽
n
(25)
Ke
Panel zone response for the initial half-cycle of loading follows the monotonic
loading rules described earlier in this paper. Thereafter, the cyclic behavior of the
panel zone is defined by elastic and inelastic curves as shown in Fig. 14. To describe
the inelastic curves, the shape factor is employed, which was first used for cyclic
stress–strain relationships by Dafalias [16]. The procedure for obtaining the shape
factor ĥ is as follows:

(i) Choose the point A such that 0.1ⱕdA / dinⱕ0.5, as shown in Fig. 15.
(ii) calculate the shape factor from

Fig. 13. Cyclic steady curve and bound line.


620 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 14. Hysteretic rules for panel zones.

Fig. 15. Shape factor for inelastic behavior.

h ⫽ dA / gAp ⫹ (din / gAp)[ln(din / dA)⫺1].


(iii) Normalize the shape factor by the plastic stiffness of the bound line ĥ ⫽
h / Kbl
p.

It has been found that a shape factor of ĥ ⫽ 20 for small rotation amplitude cycles
and ĥ ⫽ 40 for large rotation amplitude cycles provide a good correlation with
experimental data [2]. Thus, in this study a varying shape factor ĥ is used to describe
the inelastic curves of the panel zone response. The varying shape factor ĥ is updated
according to the accumulated plastic rotation qp only when unloading occurs and is
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 621

otherwise kept constant along each loading path. The varying shape factor ĥ consists
of a Boltzman function with the initial shape factor 20 (at qp ⫽ 0) and the final
shape factor 40 (at qp ⫽ ⬁) as follows:
(20⫺40)
ĥ ⫽ 40 ⫹ (26)
[1 ⫹ e(qp⫺0.213)/0.074]
It has also been found that an elastic limit factor a of 1.4 and a plastic stiffness
p ⫽ 0.008Ke provide good correlation with experimental data
of the bound line of Kbl
[2]. The position of the initial bound line is determined by drawing the line with
the slope of the bound line at the point with the corresponding slope on the cyclic
steady state curve and by making the resulting line intersect the moment axis, as
shown in Fig. 13. The plastic stiffness KAp at the point A as shown in Fig. 15 is
calculated by using the shape factor ĥ and the plastic stiffness of the bound line
Kbl
p , as follows:

KAp ⫽ Kbl 冋
p 1 ⫹ ĥ
dA
din⫺dA 册 (27)

The corresponding tangent stiffness KAt is determined by using the elastic stiffness
Ke and the plastic stiffness KAp as follows:
KeKAp
KAt ⫽ (28)
Ke ⫹ KAp
The bounding line is updated whenever load reversals occur. The procedure for
shifting the bounding line is presented below.

(i) Whenever unloading occurs, the mean values and the amplitude for the last
half cycle of the loading history, as shown in Fig. 16, are calculated.
m ⫽ 0.5(M A ⫹ M B )
M pa pa pa
(29a)
g pa
m ⫽ 0.5(g pa
A ⫹g pa
B ) (29b)
a ⫽ 0.5兩M A ⫺M B 兩
M pa pa pa
(30a)
g pa
a ⫽ 0.5兩g A⫺g
pa pa
B 兩 (30b)
where the subscripts ‘m’ and ‘a’ stand for a mean value and an amplitude, respect-
ively.
(ii) Calculate the difference between the moment amplitude Mpa a and the
moment Mpa s on the cyclic steady curve corresponding to the rotation ampli-
tude, gpa
a

⌬Mpa ⫽ Mpa
s ⫺Ma
pa
(31)
(ii) If ⌬Mpa ⬎ 0, cyclic hardening is predicted to take place in the next excur-
sion. Update the bound line by moving it outward by an amount equal to
2FH(⌬Mpa / Mpa
n ), where FH is the hardening factor.
622 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 16. Movement of bound line.

冉 冊 冉 冊
Mpa
bl

Mpa
n new

Mpa
bl

Mpa
n
⫺2FH
old
冉 冊
⌬Mpa
Mpa
n
(32)

(iv) If ⌬Mpa ⬍ 0, cyclic softening is predicted to take place in the next excursion.
Update the bound line by moving it inward by an amount equal to
2FS(⌬Mpa / Mpa
n ), where Fs is the softening factor.

冉 冊 冉 冊
Mpa
bl

Mpa
n new

Mpa
bl

Mpa
n
⫺2FS
old
冉 冊
⌬Mpa
Mpa
n
(33)

(v) Further move the bound by an amount equal to FRMpa


m , where FR is the mean
value relaxation factor.

冉 冊 冉 冊
Mpa
bl

Mpa
n new

Mpa
bl

Mpa
n
⫺FRMpa
old
m (34)

The same values used in Cofie’s study, FH ⫽ 0.45, FS ⫽ 0.07, and FR ⫽ 0.05,
are adopted in the proposed model.

6. Comparison of cyclic loading model with test data

The panel zone model for cyclic loading described above is compared with test
results for ten specimens tested by Krawinkler et al. [4], Slutter [5], Popov et al.
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 623

[18], and Engelhardt et al. [7]. For some specimens a doubler plate is used to increase
the capacity of the panel zone. Test results by Becker [3] showed that for every load
level, except maximum load, the strain in the doubler plate was significantly less
than that in the column web. Thus, the doubler plates were not fully effective. To
account for the limited participation of a doubler plate in resisting panel shear, a
reduction factor was considered in calculating the yield moments and stiffness values
of the panel zone with a doubler plate. The effectiveness of doubler plates is affected
by the method used to connect them to the column (one side attachment, both sides
attachment, welding details, etc.). In this paper, the case with a doubler plate attached
to only one side of the panel zone is studied.
Figs. 17–22 show comparisons of the analytical response predicted by the pro-
posed model and test results for panel zones with no doubler plate. These test data
include specimens with a column flange thickness up to 5.3 cm. In Figs. 20–22, the
test results for Popov et al.’s specimen 6, Krawinkler et al.’s specimen A1, and
Engelhardt et al.’s specimen DBWP are plotted against the model predictions. The
match is good for the cycles in which large deformations are imposed. For the first
few cycles in which small deformations are imposed, the predictions are not as good,
but still reasonable. For Krawinkler et al.’s specimens A2 and B2 and Slutter’s speci-
men 1 (Figs. 17–19), a large displacement amplitude was applied for the first half
cycle of loading, causing large plastic deformations, far beyond the onset of strain
hardening, in the panel zone. The model appears to work better for a large displace-
ment amplitude for which strain hardening effects are fully developed than for a
small displacement amplitude. Nonetheless, in spite of the simplicity of the model,
reasonable agreement has been achieved between model predictions and test results
for these six specimens without doubler plates.
In Figs. 23–25, model predictions are compared with test results for specimens
with doubler plates. In these specimens, the yield stress of the doubler plates was

Fig. 17. Comparison of the hysteretic rules and test data for Krawinkler et al.’s [4] specimen A2.
624 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 18. Comparison of the hysteretic rules and test data for Krawinkler et al.’s [4] specimen B2.

Fig. 19. Comparison of the hysteretic rules and test data for Slutter’s [5] specimen 1.

approximately the same as that of the column web. In the analyses, a reduction factor
of Rf ⫽ 0.4 was used to account for strain incompatibility between the column web
and the doubler plate. As indicated by these figures, reasonable agreement was achi-
eved between the model and the test data. The fact that a reduction factor of Rf ⫽
0.4 resulted in a good match with the experimental data suggests that the effective-
ness of doubler plates for increasing panel zone strength can be quite limited.
Fig. 26 shows a comparison of the analytical and experimental results for Popov
et al.’s specimen 8. For this specimen, the reported yield stress of the doubler plate
was 338 MPa, and for the column web was 413 MPa. Since the yield stress of the
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 625

Fig. 20. Comparison of the hysteretic rules and test data for Popov et al.’s [18] specimen 6.

Fig. 21. Comparison of the hysteretic rules and test data for Krawinkler et al.’s [4] specimen A1.

doubler plate was significantly different from that of the column web, two panel
elements were employed in parallel to obtain the analytical results. To obtain better
correlation between the analytical results and the test data, a reduction factor of
Rf ⫽ 0.10 was used. Once again, this suggests very limited effectiveness of the
doubler plate.
Figs. 17–26 above presented model predictions and test results for local panel
zone response for a number of test specimens reported in the literature. Fig. 27
shows a comparison between model predictions and test results for the overall load–
displacement response for Krawinkler et al.’s specimen A1. The panel zone model
626 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 22. Comparison of the hysteretic rules and test data for Engelhardt et al.’s [7] specimen DBWP.

Fig. 23. Comparison of the hysteretic rules and test data for Popov et al.’s [18] specimen 2.

described above was combined with a beam–column element by Kim and Engelhardt
[19]. Results are also compared with the simpler bilinear panel zone model. The
newly proposed panel zone model clearly provides significantly better correlation
with the test data as compared with the bilinear model.

7. Modification of cyclic loading model for composite floor slabs

The panel zone model discussed above was developed for the case where no com-
posite floor slab is present. Likewise, the comparisons between the model and experi-
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 627

Fig. 24. Comparison of the hysteretic rules and test data for Popov et al.’s [18] specimen 3.

Fig. 25. Comparison of the hysteretic rules and test data for Popov et al.’s [18] specimen 4.

mental data were all for bare steel specimens, i.e., specimens in which no composite
floor slab was present. However, since most steel moment frames have composite
floor slabs, the effects of the slab on panel zone response are of interest.
To investigate the effect of composite slabs on the behavior of panel zones, a
composite panel zone model was developed for the monotonic and cyclic behavior
of beam-to-column joints in steel moment frames with composite floor slabs. The
composite panel zone model is the same as the bare steel panel zone model described
above, except that the effective depth of the panel zone, and therefore the area of
the panel zone, is increased due to the presence of the concrete slab. This, in turn,
628 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 26. Comparison of the hysteretic rules and test data for Popov et al.’s [18] specimen 8.

Fig. 27. Comparison of test and the analysis using the developed panel zone model for overall response
for Krawinkler et al.’s [4] specimen A1.

affects the stiffness, yield moments, and cyclic steady curves of the panel zone
model. For the bare steel case, a statically equivalent force couple replaces the
moment in the beam, with the forces assumed to be acting at mid-depth of the beam
flanges. When a slab is present, and positive moment is applied to the beam, the
effective location of the force resultant at the beam’s top flange will move up towards
the slab, thereby increasing the effective panel zone depth. The bending moment
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 629

applied by the composite beam onto the panel zone will also increase for positive
bending.
The boundary forces on an interior joint panel, shown in Fig. 28, can be transfor-
med into an approximate equivalent shear force from equilibrium as follows:

Veq ⫽
M+b M⫺
+
db
⫹ ⫺
db
b
⫺ 冉
Vct ⫹ Vcb M+b M⫺
2
⬇ +
db
⫹ 冊

db
b

M+b ⫹ M⫺
Hc
b 1
⫽ +⫹ ⫺
1
d b db 冉

1
Hc冊 冉 M⫺ M+b 1⫺(d+b⫺d⫺
Mpa⫺ + ⫹ ⫺ ⬇
b

db db 冊 db⫺
+ ⫺
b ) / (2d b )⫺db / Hc pa
M (35)

1⫺(d+b⫺d⫺ +
b ) / (2db )⫺r pa
⫽ M
db⫺tbf
where db⫹ ⫽ dcom⫺ts / 2⫺tbf / 2, dcom is the composite beam depth, ts is the solid slab
thickness, and d⫺b ⫽ db⫺tbf. If db

is equal to d⫺
b then Eq. (35) reduces to Eq. (5)
for the bare steel beam-to-column joint. For a composite beam-to-column exterior
joint, Eq. (35) can be reduced to
1⫺d+b / Hc pa
Veq ⫽ M for positive moment (36a)
d+b
1⫺d⫺
b / Hc pa
Veq ⫽ M for negative moment (36b)
d⫺
b

The hysteretic rules for the composite panel zone element are based on those for
the bare steel element and are shown in Fig. 29. For the interior composite joint,
the cyclic behavior of panel zone follows path O–B–E–H, and for the exterior com-
posite joint it follows the path O–B–E⬘–H because the cyclic behavior of the exterior
joint depends on the direction of moment. To account for cracking of the concrete
slab under negative moment the unloading behavior follows the inelastic curves of

Fig. 28. Boundary forces on composite beam-to-column interior joint.


630 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 29. Hysteretic rules for composite beam-to-column joint.

the bare steel beam-to-column joint. The cyclic behavior of the panel zone under
positive moment after crack closing in the concrete slab is described by a linear
crack closing segment followed by a nonlinear composite cyclic curve (paths C–E
and F–H in Fig. 29). The factors f and b shown in Fig. 29, were determined empiri-
cally to be f ⫽ 0.5 and b ⫽ 0.1. In Fig. 30 the cyclic steady state curves for a

Fig. 30. Cyclic steady state curve for composite beam-to-column joint.
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 631

Fig. 31. Comparison of the hysteretic rules and test data for Lee’s [6] specimen EJ-FC.

composite beam-to-column joint tested by Lee [6] are shown. From this figure it can
be seen that due to the presence of a composite slab, behavior of the panel zone
differs significantly under positive and negative moments.
The proposed composite panel zone model is compared with test results for four
specimens tested by Lee [6] and Engelhardt et al. [7,8]. Also shown in these figures
are the analytical predictions using the bare steel panel zone model. Figs. 31–33
show these comparisons for specimens with no doubler plate. Specimens tested by
Lee [6] are shown in Figs. 31 and 32. The beams in these specimens were
W18 × 35 sections. For these specimens, the composite panel zone model shows

Fig. 32. Comparison of the hysteretic rules and test data for Lee’s [6] specimen IJ-FC.
632 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

Fig. 33. Comparison of the hysteretic rules and test data for Engelhardt et al.’s [7] specimen DBWP-
C.

better correlation with the test data than the bare steel panel zone model for both
small and large amplitudes of deformations. It is also clear from Figs. 31 and 32
that there is a significant difference in predicted response between the bare steel and
composite panel zone elements.
From the comparison of the model predictions and the experimental results, it can
be seen that the effects of the concrete slab are negligible for Engelhardt et al.’s
specimen DBWP-C (Fig. 33). This specimen was constructed using W36 × 150
beams. The similarity between the bare steel and composite model predictions may
be attributed to the fact that the increase of panel zone area due to the composite
slab was not significant because the steel beam depth was much larger than the slab
depth. The increase in panel zone area obtained by using Eq. (35) instead of Eq. (5)
is about 36, 16, and 8% for specimens EJ-FC (Fig. 31), IJ-FC (Fig. 32), and DBWP-
C (Fig. 33), respectively.
In Fig. 34, model predictions are compared with test results for Engelhardt and
Venti’s [8] specimen UTA-FF. This specimen was constructed with a doubler plate.
In the model, a reduction factor of Rf ⫽ 0.60 was used to account for strain incom-
patibility between the column web and the doubler plate. The performance of the
model provides reasonable correlation with experimental data. Further, there is once
again little effect of the composite slab due to the rather deep W36 × 150 beams
used for this specimen.
Only very limited experimental data are available on composite specimens with
weak panel zones. To refine mathematical models for composite panel zone response,
further experimental data are needed for specimens with composite slabs under large
inelastic excursions of monotonic loading as well as cyclic loading.
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 633

Fig. 34. Comparison of the hysteretic rules and test data for Engelhardt and Venti’s [8] specimen UTA-
FF.

8. Conclusions

The objective of the study presented in this paper was to develop models to
describe the monotonic and cyclic loading behavior of panel zones in steel moment
resisting frames. These new models build on previously developed models, and intro-
duce a number of features and refinements that show better correlation with available
experimental data.
The model for monotonic loading is based on quadri-linear panel zone moment–
deformation relations. In this model, both bending and shear deformation modes are
considered. The model proposed for cyclic loading is based on Dafalias’ bounding
surface theory combined with Cofie’s rules for movement of the bound line. Finally,
additional modifications are suggested for the cyclic loading model to account for
the influence of a composite floor slab on panel zone response. For all of the proposed
models, extensive comparisons with experimental data were presented to demonstrate
the capabilities and limitations of the models. In spite of the simplicity of the pro-
posed models, reasonable agreement between model predictions and test data was
achieved over a broad range of experiments.
In the process of developing these panel zone models, several issues were ident-
ified which appear to merit further research. One of these issues is the effect of very
thick column flanges on panel zone strength. Much of the available experimental
data are for panel zones in columns with a flange thickness less than about 3–4 cm.
In actual practice, columns with flange thickness values in excess of 10 cm are not
uncommon. Further experimental data is needed to better quantify the contribution
of very thick column flanges to overall panel zone strength.
An additional issue of concern is the effectiveness of doubler plates. Based on
comparisons between model predictions and experimental data, it appears that
634 K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635

doubler plates are, in many cases, not fully effective in contributing to panel zone
strength and stiffness. In many past experiments, it appears that doubler plates were
less than 50% effective. Additional studies are needed to better quantify the contri-
bution of doubler plates to panel zone strength and stiffness for various attachment
details. This is an issue of considerable practical importance, since doubler plates
are commonly used in practice to augment panel zone strength. Current building code
regulations in the US appear to assume that doubler plates will be fully effective.
Further studies are also needed to better quantify the effects of a composite floor
slab on panel zone behavior. The majority of past tests and past panel zone models
have considered the case of a bare steel frame. However, limited data suggests that
the presence of a composite concrete floor slab can significantly affect panel zone
behavior, particularly for relatively shallow beams. Some simple modifications to
the panel zone model were suggested in this paper to account for composite slab
effects. However, further experimental data are needed to better understand slab
effects and to further refine panel zone response models.
Finally, the models presented in this paper are intended to represent the effects
of inelasticity on panel zone response, i.e. the models account for the effects of
material yielding and strain hardening. However, after yielding and strain hardening,
the panel zone will ultimately degrade in strength either due to shear buckling of
the panel zone or due to fracture of the column or beam flanges at the corners of
the panel zone. The current models do not predict strength degradation due to insta-
bility or fracture. Further studies are needed to better understand the ultimate failure
modes of the panel zone and to quantify the available deformation capacity.

Acknowledgements

The writers gratefully acknowledge support for this work from the National
Science Foundation (grant no. CMS-9358186) and from the American Institute of
Steel Construction, Inc.

References

[1] Tsai KC, Popov EP. Steel beam–column joints in seismic moment resisting frames. Report no. EERC
88/19, University of California, Berkeley, CA; 1988.
[2] Kim K, Engelhardt, MD. Development of analytical models for earthquake analysis of steel moment
frames. Report no. PMFSEL 95-2, Phil M. Ferguson Structural Engineering Laboratory, The Univer-
sity of Texas at Austin, Austin, TX; 1995.
[3] Becker ER. Panel zone effect on the strength and stiffness of rigid steel frames. Research report,
Mechanical Lab., University of Southern California; 1971.
[4] Krawinkler H, Bertero VV, Popov EP. Inelastic behavior of steel beam-to-column subassemblages.
Report No. EERC 71/07, University of California, Berkeley, CA; 1971.
[5] Slutter RG. Tests of panel zone behavior in beam–column connections. Report no. 403.1, Fritz
Engineering Lab., Lehigh University, Bethlehem, PA; 1981.
[6] Lee SJ. Seismic behavior of steel building structures with composite Slabs. PhD thesis, Department
of Civil Engineering, Lehigh University, Bethlehem, PA; 1987.
K.D. Kim, M.D. Engelhardt / Journal of Constructional Steel Research 58 (2002) 605–635 635

[7] Engelhardt MD, Fry GT, Jones S, Venti M, Holliday S. Behavior and design of radius-cut reduced
beam section connections. Report no. SAC/BD-00/17, SAC Joint Venture, Sacramento, CA; 2000.
[8] Engelhardt MD, Venti MJ. Test of a free flange connection with a composite floor slab. Report no.
SAC/BD-00/18, SAC Joint Venture, Sacramento, CA; 2000.
[9] International Conference of Building Officials. Uniform building code, ICBO, Whittier, CA; 1997.
[10] American Institute of Steel Construction. Seismic provisions for structural steel buildings. Chicago,
IL; 1997.
[11] Federal Emergency Management Agency. Recommended seismic design criteria for new steel
moment-frame buildings. Report no. FEMA-350; 2000.
[12] Lui EM. Effects of connection flexibility and panel zone deformation on the behavior of panel steel
frames. PhD thesis, Department of Civil Engineering, Purdue University, West Lafayette, IN; 1985.
[13] Fielding DJ, Huang JS. Shear in steel beam-to-column connections. Welding J. 1971;50(7):313s–
26 (research supplement).
[14] Wang SJ. Seismic response of steel building frames with inelastic joint deformation. PhD thesis,
Department of Civil Engineering, Lehigh University, Bethlehem, PA; 1988.
[15] Kanaan AE, Powell GH. DRAIN-2D—a general purpose computer program for dynamic analysis
of inelastic plane structures with user’s guide. Report no. EERC 73/6 and 73/22, University of
California, Berkeley, CA; 1973.
[16] Dafalias YF. On cyclic and anisotropic plasticity: (I) A general model including material behavior
under stress reversals. (II) Anisotropic hardening for initially orthotropic materials. PhD thesis,
Department of Civil Engineering, University of California, Berkeley, CA; 1975.
[17] Cofie NG, Krawinkler H. Uniaxial cyclic stress–strain behavior of structural steel. J. Engng. Mech.,
ASCE 1985;111(9):1105–20.
[18] Popov EP, Amin NR, Louie JC, Stephen RM. Cyclic behavior of large beam–column assemblies.
Earthquake Spectra 1985;1(2):201–37.
[19] Kim K, Engelhardt MD. Beam–column element for nonlinear seismic analysis of steel frames. J.
Struct. Engng, ASCE 2000;126(8):916–25.

You might also like