You are on page 1of 14

EFFECT OF COLUMN STIFFENER DETAILING AND WELD

FRACTURE TOUGHNESS ON THE PERFORMANCE OF WELDED


MOMENT CONNECTIONS

R.J. Dexter, University of Minnesota, United States of America


J.F. Hajjar, University of Minnesota, United States of America
D. Lee, Research Institute of Industrial Science & Technology, South Korea

ABSTRACT
Full-scale cyclically loaded cruciform experiments with weak panel zones and
welded unreinforced flange-welded web (WUF-W) prequalified moment
connections performed well provided that the weld metal has minimum Charpy
V-Notch (CVN) toughness. Welds with low CVN and brittle fracture was
obtained in one specimen despite using weld metal certified to meet minimum
CVN requirements. Unstiffened columns perform well and alternative stiffening
details all performed well, indicating that variations in column stiffening details
did not affect the potential for fracture or low-cycle fatigue. Low-cycle fatigue
performance is compared to strain range-cycles curves extrapolated from high-
cycle S-N curves.

INTRODUCTION

Following the Northridge earthquake of January 17, 1994, extensive damage to steel
moment connections was reported (1-3). This damage most often consisted of brittle
fractures of the bottom girder flange-to-column flange Complete Joint Penetration (CJP)
groove welds. The fractures were caused by the use of low toughness welds combined with
a number of other connection detailing and construction practices that were typical prior to
the earthquake (3-5). Additionally, column stiffening practices have been cited as a possible
contributor to the fractures, largely as a result of observations that many of the connections
fractured during the Northridge earthquake lacked continuity plates and that some had weak
panel zones (6). Finite element analyses (7-10) also have shown an increase in stress and
strain concentrations in the girder flange-to-column flange CJP welds associated with
excessively weak panel zones or insufficient continuity plates. It is presumed that these
stress and strain concentrations increase the potential for fracture. As a result of these
observations, there has subsequently been a tendency to be more conservative than
necessary in designing and detailing of the continuity plates and doubler plates in steel
moment connections.

Recommendations for the seismic design of new steel moment-frame buildings (3) provide
equations for determining whether continuity plates are required, and indicate that any
required continuity plates must be of equal thickness to the girder flange for interior
connections (thinner continuity plates are permitted for exterior connections), unless
connection qualification testing demonstrates that the continuity plates are not required.
Furthermore, the connection of the continuity plates to the column flanges must be made
with CJP welds, and reinforcing fillet welds should be placed under the backing bars.

Design criteria for the limit states related to column stiffening are presented in the AISC

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 177


LRFD Specification (11). The limit states of primary importance for stiffening of connections
include Local Flange Bending (LFB), Local Web Yielding (LWY), and Panel Zone yielding
(PZ). Additional provisions for seismic design of doubler plates and continuity plates were
included in the AISC Seismic Provisions (12). However, AISC (13) removed all continuity
plate design procedures for Intermediate and Special Moment Frames, requiring instead that
they be proportioned based on connection qualification tests.

The tendency towards being more conservative than necessary in column stiffener design
has raised concerns about economy as well as the potential for cracking of the k-area in the
column web near the web-flange junction during fabrication due to high residual stresses
caused by highly restrained CJP welds on the continuity plates or doubler plates (14,15).
Therefore, a combined experimental and computational research study was conducted at the
University of Minnesota to reassess the recent column stiffener design and detailing
provisions and recommendations, and to provide economical alternative stiffener details that
minimize welding along the column k-line while retaining superior performance for non-
seismic and seismic design (16). This paper examines the effect of variations in column
stiffening, including no stiffening as well as more economnical alternative stiffening details,
on the fracture and low-cycle fatigue performance of Welded Unreinforced Flange-Welded
Web (WUF-W) moment connections (3).

The design and results of these tests are reported elsewhere, including an assessment of
LFB and LWY limit states (17) and the cyclic panel zone behavior and design (18). Related
research included nine pull-plate experiments (19-22) that investigated the limit states of LFB
and LWY, primarily for non-seismic design, and further tested the alternative doubler plate
and continuity plate stiffener details. Finite element analyses of all experimental specimens
were also conducted as part of this research as well as parametric studies to extend the
results to member sizes and details not tested (23).

FULL-SCALE CONNECTION TESTS

Six full-scale, girder-to-column cruciform specimens were tested (Table 1). The SAC (24)
loading history was applied including six cycles at each interstory drift level of 0.375%, 0.5%,
and 0.75%, four cycles at 1.0% interstory drift level, and two cycles at each interstory drift
level of 1.5%, 2.0%, 3.0%, and 4.0%.

ASTM A992 wide-flange sections and A572 Grade 50 plate was used. The specimens used
the pre-qualified (3) WUF-W connection detail (Figure 1). The column stiffening was varied
in these specimens including three alternative doubler plate details (i.e., back-beveled fillet-
welded doubler plate, square-cut fillet-welded doubler plate, and groove-welded box doubler
plate; see Figure 2) and a fillet-welded ½ in. thick fillet-welded continuity plate detail.

Table 2 presents the design strength-to-demand ratios using minimum specified material
properties for the LFB and LWY limit states. Note that the design strength for LFB and LWY
is the design strength of the column shape alone and does not include the column
reinforcement, if any. The demand is calculated with various methods:

• Yield strength of the girder flange,


• A value of 1.8 times the yield strength of the flange as in the 1992 AISC Seismic
Provisions (12)
• Equation included in AISC Design Guide No. 13 (25) which amplifies the plastic
moment due to shear and reduces this moment to a force couple of the flanges. This
equation, or slightly modified forms of it, has been widely used for the design of

178 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


column stiffeners in steel moment connections (3).

Table 2 shows that 1992 AISC Seismic Provisions (12) and AISC Design Guide No. 13 (25)
provide similar demand values. For these cases, the demand exceeds the capacity for LFB
for all specimens except CR1, whereas only specimen CR3 has continuity plates (although
the box detail functions as a continuity plate as well as a doubler). The remaining specimens
CR2 and CR5 are underdesigned for LFB for seismic demand.

Table 1. Test matrix of cruciform specimens.


CR4
CR1 CR2 CR3 CR5
and CR4R
Girder W24x94 W24x94 W24x94 W24x94 W24x94
Column W14x283 W14x193 W14x176 W14x176 W14x145
Doubler Plate Detail III
None Detail II Detail II Detail I
(DP) Box (Offset)
DP Thickness NA 0.625 in. 2 @ 0.5 in. 2 @ 0.75 in. 2 @ 0.625 in.
Continuity Fillet-
None None None None
Plate (CP) welded
CP Thickness NA NA 0.5 in. NA NA

Figure 1. Typical welding details used for cruciform specimens (Specimen CR1).

However, the girder flange demand predicted by the latter two methods is very conservative
and can be put in perspective by comparing to the maximum possible uniaxial tensile
strength of A992 steel. The stress in the flange is 1.8 times 50 ksi or 90 ksi, well above the
likely tensile strength of A992 steel. For example, a survey of more than 20,000 mill reports
from (26,27) showed that A992 steel has a mean tensile strength of 73 ksi. The 97.5
percentile tensile strength was 80 ksi, and the maximum value reported was 88 ksi.

Also shown in Table 2 are the panel zone capacity-to-demand ratios (including the strength
of the doubler plate in the capacity) and column-girder moment ratios calculated from the

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 179


2002 AISC Seismic Provisions (28), assuming no axial compression in the column. The
panel zone demand exceeds the capacity for all specimens. Therefore, if column stiffening
were necessary to prevent premature brittle fracture or low-cycle fatigue, these specimens
are a worst-case test since they are underdesigned.

Figure 2. Doubler details: (a) back-beveled fillet-welded doubler (Detail I), (b) square-cut
fillet-welded doubler (Detail II), (c) box (offset) doubler (Detail III).

Table 2. Nominal capacity/demand ratios of PZ yielding, LFB, and LWY limit states.
LFB φRn/Ru LWY φRn/Ru
∑ M *pc PZ
∑ M *pb φvRv/Ru 1.8Yield a 1.8Yield a
a a a
(AISC) Yield 1992 DG13 Yield 1992 DG13 a
(AISC)
Seismic Seismic
3.04 1.69 1.64
CR1 1.50 0.72 2.38 1.32 1.29
(3.38)b (1.88) (1.82)
1.47 0.82 0.80
CR2 0.99 0.66 2.20 1.22 1.19
(1.63) (0.91) (0.89)
1.22 0.68 0.66
CR3 0.89 0.74 2.51 1.39 1.36
(1.36) (0.76) (0.73)
CR4 1.22 0.68 0.66
0.89 0.93 3.19 1.77 1.73
CR4R (1.36) (0.76) (0.73)
0.84 0.47 0.46
CR5 0.73 0.74 2.34 1.30 1.27
(0.93) (0.52) (0.51)
a
Equation used to calculate demand, Ru
b
Values in parentheses reflect use of φ = 1.0

Weld details

E70T-1 (Lincoln Outershield 70) wire with 100% CO2 shielding gas was used for all shop
welding. The girder flange-to-column flange CJP groove welds were made in the flat position
with E70T-6 (Lincoln Innershield NR-305) wire. Welds made with E70T-6 wire are required
by AWS A5.20 (29) and AISC 2002 Seismic Specifications (28) to have notch toughness of
20 ft-lbs at -20°F. FEMA 350 (3) has recommended minimum notch toughness requirements
at two temperatures, 20 ft-lbs at 0°F and 40 ft-lbs at 70°F. According to the Lincoln Electric
Company product family literature, the typical values for NR-305 are 21 to 35 ft-lbs at -20°F
and 21 to 54 ft-lbs at 0°F. As shown in Table 3, Specimens CR1 and CR4 were fabricated
with a 5/64 in. diameter NR-305 wire and the remaining were fabricated with 3/32 in.
diameter NR-305 wire. All CJP welds were ultrasonically tested by a certified inspector in
conformance with Table 6.3 of AWS D1.1-2000 (30) for cyclically loaded joints.

The out-of-position field welds, including the CJP welds connecting the girder web to the
column flange and all reinforcing fillet welds were made with 0.068 in. diameter E71T-8
(Lincoln Innershield NR-203MP) wire for Specimens CR1 and CR4, and 5/64 in. diameter

180 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


E71T-8 (Lincoln Innershield NR-232) wire for the other specimens.

The shear tab was designed to extend approximately 0.25 in. into the top and bottom access
holes and acted as the backing bar for the CJP welds of the girder web to the column flange.
This extension acted as a short runoff tab, allowing the weld to extend the full depth of the
girder web. Ricles et al. (31) recommended that these runoff tabs of the vertical web weld be
ground smooth, which is labor intensive. Since it was felt that this might not be necessary,
these runoff tabs were not ground smooth in the specimens tested in this work.

Table 3. Tested weld material properties (E70T-6 only).


E70T-6a E70T-6
5/64 in. wire 3/32 in. wire
CR1a CR4a CR2 CR3 CR4R CR5
CVN @ 0°F
2.6 2.0 34.3 44.3 33.0 33.0
(ft-lbs)
CVN @ 70°F
19.3 2.3 54.3 73.3 58.7 53.7
(ft-lbs)
Fy (ksi) NA NA 59.5 50.0 56.0 53.5
Fu (ksi) NA NA 79.5 72.5 78.2 75.5
% Elongation NA NA 25.0 23.0 27.5 26.0
a
For Specimens CR1 and CR4, the CVN tests were performed on specimens machined after
the experiment from the welds that did not fracture in the cruciform joints

BRITTLE FRACTURE

Specimen CR4 exhibited premature brittle failure in three of four girder flange-to-column
flange CJP welds in the early stage of the SAC (1997) loading history and was stopped after
one-half cycle at 2.0% interstory drift. It was found that this specimen was unintentionally
prepared with low-toughness weld metal, as shown in Table 3. Note that the AWS
Certificate of Conformance for this wire indicated that the weld metal meets the minimum
toughness requirement of AWS A5.20 (28) of 20 ft-lbs at -20°F (32).

Specimen CR4R was essentially a replicate test of Specimen CR4, except that the weld
metal used for Specimen CR4R met the minimum requirements of FEMA 350 (3). Specimen
CR4R not only performed acceptably according to the SAC (24) requirements, it performed
as well as any of the other specimens successfully tested in this experimental study. The
fact that the box (offset) doubler plate detail performed well in Specimen CR4R indicates that
the detail itself was probably not a factor in the fracture that occurred in Specimen CR4.

Following the premature brittle failure in Specimen CR4, it was found that the previously
tested Specimen CR1 also had relatively low weld-metal notch toughness, with an average
of 2.6 ft-lbs at 0°F and 19.3 ft-lbs at 70°F as presented in Table 3. Specimen CR1, which
was welded using the same wire and the similar welding procedure as Specimen CR4, but
had marginally better notch toughness, performed very well, experiencing 14 cycles of 4.0%
interstory drift before the significant strength degradation. It is important to note that this
Specimen CR1 had no doubler plates or continuity plates, even though doubler plates would
be required as shown in Table 2. Thus this test shows that column stiffeners are not
absolutely required to avoid brittle fractire or low-cycle fatigue, even with this very poor notch
toughness. These two tests have closely bracketed the minimum notch toughness required
for adequate performance of CJP welds.

It is believed that the FEMA 350 requirements (3) for minimum notch toughness are

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 181


adequate, provided they can be consistently met. FEMA 353 (4) requires toughness testing
on each production lot of the specified filler metal. However, lot testing is typically not done
since FEMA 353 also allows this requirement to be waived (upon approval of the Engineer),
instead relying upon the consumable manufacturer’s certification. However, the certification
is not necessarily reliable; for example, this 5/64-in. diameter E70T-6 wire that produced
these brittle welds had been certified by the manufacturer as meeting the minimum 20 ft-lbs
at -20°F required by the AWS certification test (32). Therefore, further evaluation of the CVN
testing requirements for weld metal is warranted.

LOW-CYCLE FATIGUE

The remaining connections exhibited no brittle fractures, but low-cycle fatigue failures occurred
after significant cyclic loading. Figures 3 and 4 show low-cycle fatigue cracks forming at the
beam flange weld. These beam flange weld cracks were the only type of low-cycle fatigue
cracks that actually propagated to cause failure, which is defined here as significant strength
reduction. Low-cycle fatigue cracks did originate at the weld of the beam web to the column
flange and at the weld access hole, as shown in Figure 5. However after propagating for a short
distance they arrested and did not propagate further or lead to failure, so therefore they are not
structurally significant.

(a) (b)
Figure 3. Low cycle fatigue crack developing at the toe of the beam flange weld in a moment-
frame connection after (a) 11 cycles and (b) 17 cycles of 4% drift.

Table 4 shows the cycles at 3% or 4% drift when the first crack was first visible in the CJP
welds, the final cycles at 4% drift when the strength was reduced, and the measured strain
ranges. The pairs of measured strain ranges for each specimen are from the west top flange
and the east bottom flange, respectively (except for CR1 where only the west top flange data
were avaialble). The average of five gages across the width was used to eliminate some of
the scatter and the effect of strain gradients. It is believed that the variation in the measured
strain ranges is random, and that the strain range at 4% drift was relatively consistent among
the specimens, averaging approximately 4.1%.

The performance requirement is that the connections must complete 2 cycles at 4% drift without
a significant reduction in strength in order to be prequalified connections (3). One could
conclude that all these connections (except the original CR4, which experienced brittle fracture)
met this performance requirement and that therefore the performances of the specimens are
equally good. However, there may be some significance to the final number of cycles before

182 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


strength reduction. In particular, it is believed that the differences in number of cycles to strength
reduction between 12 and 16 are due to random variability, and that therefore specimens CR1,
CR2, CR3, and CR4R performed equally well. This means that the variation in column stiffener
detailing and panel zone strength among these specimens had no significant influence on the
low-cycle fatigue performance. For specimen CR5, the number of cycles before strength
reduction is 6. As shown in Table 2, this specimen had the lowest ratio of capacity to demand
for LFB – less than one even for non-seismic (nominal yield strength of flange) and less than 0.5
for either seismic demand. It is likely that the relatively smaller number of cycles in this under-
designed specimen is due to lack of continuity plates.

LCF Crack

Column Column
Web Flange

Slag Inclusion,
LOF

Figure 4. Cross section of beam flange weld showing low cycle fatigue crack
developing at the weld toe.

Figure 5. Low cycle fatigue cracks forming at the end of the beam web to column flange
weld and at the weld access hole.

Table 4. Low-cycle fatigue data.


Specimen CR1 CR2 CR3 CR4R CR5
Column W14x283 W14x193 W14x176 W14x176 W14x145
Doubler Plate None Detail II Detail II Box Detail I
Continuity Plate None None Fillet welded None (box) None
Cycles when Crack
11@4% 2@3% 2@3% 2@4% 1@3%
Visible
Cycles at 4% when
14 16 14 12 6
Strength Reduced
Strain Range in
Girder Flange at 4.2% 4.5%, 3.7% 3.8%, 4.5% 3.8%, 3.8% 3.7%, 4.8%
4% Drift

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 183


As noted above, specimen CR1 had much lower notch toughness than the others.
Therefore, it appears that, as long as the notch toughness is sufficient to preclude brittle
fracture, notch toughness above this minimum level also has little influence on low-cycle
fatigue performance. Finally, for Specimens CR1 through CR4, there is little correlation
betweeen the cycles when the crack was first detected in the CJP welds, which was highly
variable, and the final number of cycles when strength was reduced. Therefore, it is believed
that the number of cycles at which the crack is first detected is not significant.

Most past research on low-cycle fatigue has involved pressure vessels and some other types
of mechanical engineering structures. Since low-cycle fatigue is an inelastic phenomenon,
the strain range is the key parameter rather than the stress range as in high-cycle fatigue.
The Coffin-Manson rule (33) has been used to relate the strain range in smooth tensile
specimens to the fatigue life. Manson suggested a conservative lower-bound simplification,
called Manson’s universal slopes equation (34):
σu
∆ε = 3.5 N −0.12 + ε 0f .6 N −0.6 (1)
E
where: ∆ε is the total strain range, σu is the tensile strength, and εf is the elongation at fracture.

Note that the first term in Equation 1 is the elastic part of the total strain range (which is relatively
insignificant when there are fewer than 100 cycles) and the second term is the plastic part of the
total strain range. Figure 6 shows a plot of Manson’s universal slopes equation where σu is 450
MPa and εf is 25%, typical minimum properties for Grade 50 structural steel. Many studies have
shown that Manson’s universal slopes equation is conservative compared to experimental data
from smooth specimens (34,35). However, because of buckling at greater strain ranges, most of
the experimental data are for strain ranges less than 1%, i.e., for cycles greater than 1000.
Limited data exist at higher strain ranges – some are shown in Figure 6 for A36 steel smooth
specimens machined from the flanges of wide-flange sections (35).

At this time, very little is understood about low-cycle fatigue in welded or bolted structural details.
For example, it is a very difficult task to predict accurately the local strain range at a location of
cyclic local flange buckling. However, Krawinkler and Zohrei (36) and Ballio and Castiglioni
(37,38) showed that the number of cycles to failure by low-cycle fatigue of welded connections
could be predicted by the local strain range in a power law that is analogous to an S-N curve.
Ballio and Castiglioni (37,38) showed that the power law would have and exponent of 3, just like
the elastic S-N curves. Krawinkler and Zohrei (36) also showed that Miner’s rule (39) could be
used to predict the number of variable-amplitude cycles to failure based on constant amplitude
test data.

Therefore, it may be possible to predict and design against low-cycle fatigue using strain-range
vs. number-of-cycles curves that are extrapolated from the high-cycle fatigue design S-N curves.
Figure 6 shows the AISC S-N curves (11) for Categories A and C, converted from stress range
to strain range by dividing the stress ranges by the elastic modulus, and extrapolated up to one
cycle.

There are only limited data to support this approach. Figure 6 shows the strain range-number of
cycles data from these WUF-W beam-to-column connection tests. The number of cycles plotted
in Figure 6 is the number of cycles at 4% drift. If the effect of all the previous cycles is included
using Miner’s rule (39), they add up to an equivalent of about one additional cycle at 4%. Since
flange strains right at the weld toe were used rather than nominal values, this is analogous to a
hot-spot approach for high-cycle fatigue. For high-cycle fatigue, the Category C S-N curve is a
suitable baseline S-N curve for the hot-spot approach (40,41). It appears that the Category C S-
N curve is also a good lower bound to these low-cycle fatigue data. The scatter in the data is
substantial, as is also true in high-cycle fatigue.

184 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


Also shown in Figure 6 are previously unpublished data for smaller coupon-type specimens with
transverse butt welds, which would be expected to be Category C details. These are some of
the only available data with fewer than 5,000 cycles. These coupon data are also in reasonable
agreement with the extrapolated Category C curve as a lower bound.
100
welded coupon
Manson's Equation smooth specimen
WUF-W connection tests
Strain Range [%]

Category A
10
Category C

0.1
1 10 100 1000 10000
Cycles
Figure 6. Comparison of standard S-N curves presented in terms of strain range and Manson’s
universal slopes equation for Grade 50 (350 MPa yield strength) steel to low-cycle fatigue test
data and the connection test data.

CONCLUSIONS AND RECOMMENDATIONS

1. When properly detailed and welded with notch-tough filler metal, the WUF-W steel
moment connections can perform adequately even though relatively weak panel zones and
low local flange bending strengths were chosen.
2. The failure mode of the specimens other than the original CR4 was low-cycle fatigue
(LCF) crack growth and eventual rupture of one or more girder flange-to-column flange
complete joint penetration (CJP) groove welds. Low cycle fatigue may be conservatively
predicted using strain-range vs. cycles curves derived from the stress based S-N curves for
high-cycle fatigue.
3. Specimens CR1 and CR4 were unintentionally prepared with weld metal that had CVN
values that were much lower than the minimum requirements. The premature brittle failure of
specimen CR4 reconfirmed that achieving the required minimum CVN toughness in the
girder flange-to-column flange CJP welds is critical. These low toughness welds occurred
despite the certification of the filler metal; the certification is only required annually, unlike the
way that each heat of steel is tested. A study should be conducted to fully characterize the
typical variability in the CVN and other properties of the weld. An evaluation of the need for
lot testing should be performed. Consideration should also be given to use of filler metals
with a distribution of CVN such that there is a sufficiently small probability of not meeting the
minimum required values, and therefore lot testing may not be required.
4. Application of the alternative column stiffener details (i.e., back-beveled fillet-welded
doubler plate detail; square-cut fillet-welded doubler plate detail; groove-welded box doubler
plate detail; fillet-welded 1/2 in. thick continuity plates) in the WUF-W steel moment
connections was successfully verified. No cracks or distortions were observed in the welds
connecting these stiffeners before the rupturing of the girder flange-to-column flange CJP

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 185


welds.
5. Specimens CR1, CR2, CR4R, and CR5, none of which had continuity plates (although
Specimen CR4R included the offset doubler plate detail), met the requirements for two
cycles at 4.0% interstory drift, athough only Specimen CR1 met the seismic requirements of
AISC and FEMA with respect to continuity plates. Continuity plates may thus not be
necessary in many interior columns in steel moment connections, and design provisions
permitting the design, or lack of inclusion, of continuity plates are recommended for
reintroduction into the AISC Seismic Provisions.
6. For a wide range of column sections and doubler plate detailing, strain gradients and
strain magnitudes well above the yield strain in the girder flanges did not prohibit the
specimens from achieving the connection prequalification requirement of completing two
cycles at 4.0% interstory drift without significant strength degradation. This was even the
case for specimen CR1, which had notch toughness in the weld metal that was significantly
below the requirements. These results indicate that the column reinforcement detailing may
not have a significant effect on the potential for brittle fracture at the girder flange-to-column
flange weld. Note that this is contrary to previous finite-element analyses reported in the
literature using theoretical fracture criteria that have predicted a significant effect of using or
omitting continuity plates.
7. If continuity plates are required, fillet-welded continuity plates that were approximately
half of the girder flange thickness performed well. The results showed that only minor local
yielding occurred in these continuity plates in part of the most stressed cross sections at
peak drift level and that these strains were not sufficient to cause cracking or distortion in the
continuity plate or to change the strain gradients in the girder flange substantially. Since
continuity plates do not significantly yield, it may not be necessary to size the welds large
enough to develop the continuity plate. Rather the weld and the plate may only need to be
designed for the difference between the demand and the capacity of the column shape
without the continuity plate. Continuity plates with undersized fillet welds should be tested to
confirm that the weld need not develop the full continuity plate strength.
8. In all the tests except the original CR4, the seismic performance of the relatively weak
panel zones was stable and ductile, and the panel zones exhibited good energy dissipation.
Lee et al. (17-23) provide recommended changes to the AISC panel zone strength
equations, as well as detailed evaluations of current AISC provisions for local flange bending,
local web yielding, and panel zone shear.

ACKNOWLEDGEMENT

This research was sponsored by the American Institute of Steel Construction, Inc. and by the
University of Minnesota. In-kind funding and materials were provided by LeJeune Steel
Company, Minneapolis, Minnesota; Danny’s Construction Company, Minneapolis,
Minnesota; Braun Intertec, Minneapolis, Minnesota; Nucor-Yamato Steel Company,
Blytheville, Arkansas; Lincoln Electric Company, Cleveland, Ohio; and Edison Welding
Institute, Cleveland, Ohio. Supercomputing resources were provided by the Minnesota
Supercomputing Institute. The authors wish to thank T. V. Galambos and P. M. Bergson,
University of Minnesota, L. A. Kloiber, LeJeune Steel Company, and the members of the
technical advisory group on this project for their valuable assistance.

REFERENCES

1. Youssef, N. F. G, Bonowitz, D., and Gross, J. H. (1995). A Survey of Steel Moment-


Resisting Frame Buildings Affected by the 1994 Northridge Earthquake, Report No.
NISTIR 5625, National Institute of Standards and Technology, Gaithersburg, Maryland.

186 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


2. Northridge Reconnaissance Team (1996). Northridge Earthquake of January 17, 1994,
Reconnaissance Report (Supplement C-2 to Volume 11), EERI, Oakland, California.
3. Federal Emergency Management Agency (FEMA) (2000). Recommended Seismic
Design Criteria for New Steel Moment-Frame Buildings, Report No. FEMA 350, FEMA,
Washington, D.C.
4. FEMA (2000). Recommended Specifications and Quality Assurance Guidelines for
Steel Moment-Frame Construction for Seismic Applications, Report No. FEMA 353,
FEMA, Washington, D.C.
5. Fisher, J.W., R.J. Dexter, and E.J. Kaufmann, “Fracture Mechanics of Welded Structural
Steel Connections,” Report No. SAC 95-09, FEMA-288, March 1997
6. Tremblay, R., Timler, P., Bruneau, M., and Filiatrault, A. (1995). “Performance of Steel
Structures during the 1994 Northridge Earthquake,” Canadian Journal of Civil
Engineering, Vol. 22, pp. 338-360.
7. El-Tawil, S., Mikesell, T., Vidarsson, E., and Kunnath, S. (1998). “Strength and Ductility
of FR Welded-Bolted Connections,” Report No. SAC/BD-98/01, SAC Joint Venture,
Sacramento, California.
8. El-Tawil, S., Vidarsson, E., Mikesell, T., and Kunnath, S. K. (1999). “Inelastic Behavior
and Design of Steel Panel Zones,” Journal of Structural Engineering, ASCE, Vol. 125,
No. 2, pp. 183-193.
9. El-Tawil, S. (2000). “Panel Zone Yielding in Steel Moment Connections,” Engineering
Journal, AISC, Vol. 37, No. 1, pp. 120-131.
10. Mao, C., Ricles, J. M., Lu, L., and Fisher, J. W. (2001). “Effect of Local Details on
Ductility of Welded Moment Connections,” Journal of Structural Engineering, ASCE, Vol.
127, No. 9, pp. 1036-1044.
11. American Institute of Steel Construction (AISC) (1999). Load and Resistance Factor
Design Specification for Structural Steel Buildings, AISC, Chicago, Illinois.
12. AISC (1992). Seismic Provisions for Structural Steel Buildings, AISC, Chicago, Illinois.
13. AISC (1997). Seismic Provisions for Structural Steel Buildings, AISC, Chicago, Illinois.
14. AISC (1997). “AISC Advisory on Mechanical Properties Near the Fillet of Wide Flange
Shapes and Interim Recommendations, January 10, 1997,” Modern Steel Construction,
AISC, Chicago, Illinois, February, p. 18.
15. Tide, R. H. R. (2000). “Evaluation of Steel Properties and Cracking in k–area of W
Shapes,” Engineering Structures, Vol. 22, No. 2, pp. 128-134.
16. Lee, D., Cotton, S., Dexter, R. J., Hajjar, J. F., Ye, Y., and Ojard, S. D. (2002). “Column
Stiffener Detailing and Panel Zone Behavior of Steel Moment Frame Connections, ”
Report No. ST-01-3.2, Department of Civil Engineering, University of Minnesota,
Minneapolis, Minnesota.
17. Lee, D., Cotton, S. C., Hajjar, J. F., Dexter, R. J., Ye, Y., and Ojard, S. D. (2004). “Cyclic
Behavior of Steel Moment-Resisting Connections Reinforced by Alternative Column
Stiffener Details. I. Connection Performance and Continuity Plate Detailing,”
Engineering Journal, AISC, submitted for publication.
18. Lee, D., Cotton, S. C., Hajjar, J. F., Dexter, R. J., Ye, Y., Ojard, S. D. (2004). “Cyclic
Behavior of Steel Moment-Resisting Connections Reinforced by Alternative Column
Stiffener Details. II. Panel Zone Behavior and Doubler Plate Detailing,” Engineering
Journal, AISC, submitted for publication.
19. Prochnow, S. D., Dexter, R. J., Hajjar, J. F., Ye, Y., and Cotton, S. C. (2000). “Local
Flange Bending and Local Web Yielding Limit States in Steel Moment-Resisting
Connections,” Report No. ST-00-4, Department of Civil Engineering, University of
Minnesota, Minneapolis, Minnesota.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 187


20. Prochnow, S. D., Ye., Y., Dexter, R. J., Hajjar, J. F., and Cotton, S. C. (2002). “Local
Flange Bending and Local Web Yielding Limit States in Steel Moment-Resisting
Connections,” Connections in Steel Structures IV, Roanoke, Virginia, October 22-24,
2000, American Institute of Steel Construction, Chicago, Illinois, pp. 318-328.
21. Dexter, R. J., Hajjar, J. F., Prochnow, S. D., Graeser, M. D., Galambos, T. V., and
Cotton, S. C. (2001). “Evaluation of the Design Requirements for Column Stiffeners and
Doublers and the Variation in Properties of A992 Shapes,” Proceedings of the North
American Steel Construction Conference, AISC, Chicago, Illinois.
22. Hajjar, J. F., Dexter, R. J., Ojard, S. D., Ye, Y., and Cotton, S. C. (2003). “Continuity
Plate Detailing for Steel Moment-Resisting Connections,” Engineering Journal, AISC,
Vol. 40, No, 4, Fourth Quarter, pp. 189-211.
23. Ye, Y., Hajjar, J. F., Dexter, R. D., Prochnow, S. C., and Cotton, S. C. (2000).
“Nonlinear Analysis of Continuity Plate and Doubler Plate Details in Steel Moment
Frame Connections,” Report No. ST-00-3, Department of Civil Engineering, University of
Minnesota, Minneapolis, Minnesota.
24. SAC (1997). “Protocol for Fabrication, Inspection, Testing, and Documentation of Beam-
Column Connection Tests and Other Experimental Specimens,” Report No. SAC/BD-
97/02, SAC Joint Venture, Sacramento, California.
25. AISC (1999). “Stiffening of Wide-Flange Columns at Moment Connections: Wind and
Seismic Applications,” AISC Design Guide No. 13, AISC, Chicago, Illinois.
26. Dexter, R. J. (2000). “Structural Shape Material Property Survey, ” Final Report to
Structural Shape Producer's Council, University of Minnesota, Minneapolis, Minnesota.
27. Bartlett, F.M., R.J. Dexter, M.D. Graeser, J.J. Jelinek, B.J. Schmidt, and T.V. Galambos
(2003). “Updating Standard Shape Material Properties Database for Design and
Reliability”, Engineering Journal, AISC, Vol. 40, No. 1, 1st Quarter, pp 2-14.
28. AISC (2002). Seismic Provisions for Structural Steel Buildings, AISC, Chicago, Illinois.
29. American Welding Society (AWS) (1995). Specification for Carbon Steel Electrodes for
Flux Cored Arc Welding, AWS A5.20-95, AWS, Miami, Florida.
30. AWS (2000). Structural Welding Code – Steel, AWS D1.1-2000, AWS, Miami, Florida.
31. Ricles, J. M., Mao, C., Lu, L., and Fisher, J. W. (2002). “Inelastic Cyclic Testing of
Welded Unreinforced Moment Connections,” Journal of Structural Engineering, ASCE,
Vol. 128, No. 4, pp. 429-440.
32. Lincoln Electric Co. (1999). Certificate of Conformance for Innershield NR-305
(E70T-6), Lincoln Electric Co., Cleveland, Ohio, August 6, 1999.
33. Coffin, L.F. Jr., “A Note on Low Cycle Fatigue Laws”, Journal of Materials, Vol. 6, No. 2,
pp.388-402, 1971.
34. Itoh, Y.Z., Kashiwaya, H., “Low-cycle fatigue properties of steel s and their weld metals”,
Journal of Engineering Materials and Technology, Vol. 111, No. 4, pp. 431-437, 1989.
35. Howdyshell, P., J.C. Trovillion, and J.L. Wetterich, “Low-Cycle Fatigue of Structural
Materials,” Materials and Construction: Proceedings of MatCong 5, the 5th ASCE
Materials Engineering Congress, Bank, L. (ed.), 10-12 May 1999, Cincinnati, OH,
American Society of Civil Eng., Reston, VA., pp. 148-155.
36. Krawinkler, H., and Zohrei, M. (1983). “Cumulative Damage in Steel Structures
Subjected to Earthquake Ground Motion,” Computers and Structures, Vol. 16, No. 1-4,
pp.531-541, 1983.
37. Ballio, G., and Castiglioni, C.A. (1995). “A Unified Approach for the Design of Steel
Structures Under Low and/or High Cycle Fatigue,” Journal of Constructional Steel
Research, Vol. 34, No. 1, pp. 75-101
38. Castiglioni, C.A. (1995). “Cumulative Damage Assessment in Structural Steel Details,”

188 Connections in Steel Structures V - Amsterdam - June 3-4, 2004


IABSE Symposium San Francisco, Extending the Lifespan of Structures, pp. 1061-1066.
39. Miner, M., “Cumulative Damage in Fatigue,” Transactions of the American Society of
Mechanical Engineers, Vol. 67.
40. Dexter, R.J., Tarquinio, J.E., and Fisher, J.W., "Application of the hot-spot stress fatigue
analysis to attachments on flexible plate," Proceedings, 13th International Conference on
Offshore Mechanics and Arctic Engineering, ASME, Vol. III, Materials Engineering, 85-
92.
41. Dexter, R. J., “Fatigue and fracture", The Structural Engineering Handbook, 2nd Edition,
Lui, B. M. ed., CRC Press, 2004.

Connections in Steel Structures V - Amsterdam - June 3-4, 2004 189


190 Connections in Steel Structures V - Amsterdam - June 3-4, 2004

You might also like