Está en la página 1de 156

Eng ine e ring

J ourna l
AMERICAN INSTITUTE OF STEEL CONSTRUCTION, INC.
Pa g e 1: La wre nc e G. Griffis
Se rvic a b ility Limit Sta te s Und e r Wind Loa d
Pa g e 17: A. Zure ic k
De sig n Stre ng th of Conc e ntric a lly Loa d e d Sing le
Ang le Struts
Pa g e 31: Tony Lue a nd Dua ne S. Ellifritt
The Wa rp ing Consta nt for the W-Se c tion with a
Cha nne l Ca p
Pa g e 34: Cha rle s J . Ca rte r a nd Louis F.
Ge sc hwind ne r
The Ec onomic Imp a c t of Ove rsp e c ifying Simp le
Conne c tions
Pa g e 37: Disc ussionWillia m E. Moore II
Simp le Eq ua tions for Effe c tive Le ng th
Fa c torsPie rre Dumonte il
Pa g e 38: Corre c tionPie rre Dumonte il
Simp le Eq ua tions for Effe c tive Le ng th
Fa c tors
Pa g e 39: Corre c tionLe wis B. Burg e tt
Fa st Che c k for Bloc k She a r
1st Qua rte r 1993/ Volume 30, No. 1
INTRODUCTION
The increasing use and reliance on probability based limit
states design methods, such as the recently adopted AISC
LRFD Specification,
1
has focused new attention on the prob-
lems of serviceability in steel buildings. These methods,
along with the development of higher-strength steels and
concretes and the use of lighter and less rigid building mate-
rials, have led to more flexible and lightly damped structures
than ever before, making serviceability problems more
prevalent.
The purpose of this paper is to focus attention on two
important serviceability limit states under wind loads;
namely, deformation (including deflection, curvature, and
drift) and motion perception (acceleration). These issues are
particularly important for tall and/or slender steel and com-
posite structures. A brief review of available information on
these subjects will be presented followed by a discussion of
current standards of practice, particularly in the United States.
Finally, proposed standards will be presented that, hopefully,
will focus attention, debate, and perhaps new research efforts
on these very important issues in design.
IMPORTANCE OF SERVICEABILITY
LIMIT STATES
12,31
Every building or other structure must satisfy a strength limit
state, in which each member is proportioned to carry the
design loads to resist buckling, yielding, instability, fracture,
etc.; and serviceability limit states which define functional
performance and behavior under load and include such items
as deflection, vibration, and corrosion. In the United States,
strength limit states have traditionally been specified in build-
ing codes because they control the safety of the structure.
Serviceability limit states, on the other hand, are usually
noncatastrophic, define a level of quality of the structure or
element, and are a matter of judgment as to their application.
Serviceability limit states involve the perceptions and expec-
tations of the owner or user and are a contractual matter
between the owner or user and the designer and builder. It is
for these reasons, and because the benefits themselves are
often subjective and difficult to define or quantify, that ser-
viceability limit states for the most part are not included
within U.S. building codes. The fact that serviceability limit
states are usually not codified should not diminish their
importance. Exceeding a serviceability limit state in a build-
ing or other structure usually means that its function is dis-
rupted or impaired because of local minor damage, deterio-
rations, or because of occupant discomfort or annoyance.
While safety is usually not at issue, the economic conse-
quences can be substantial. Interestingly, there are some
serviceability items that can also be safety related. For in-
stance, excessive building drift can influence frame stability
because of the P- effect. Excessive building drift can also
cause portions of the building cladding to fall and potentially
injure pedestrians below.
Serviceability limit states can be grouped into three cate-
gories as follows:
1. Deformation (deflection, curvature, drift). Common ex-
amples include local damage to nonstructural elements
(e.g., ceilings, cladding, partitions) due to deflections
under dead, live, wind, or seismic load; and damage
from temperature change, moisture, shrinkage, or creep.
2. Motion perception (vibration, acceleration). Common
examples include human discomfort caused by wind or
machinery, particularly if resonance occurs. Floor vibra-
tions from people or machinery and acceleration in tall
buildings under wind load are usual areas of concern in
this category.
3. Deterioration. Included are such items as corrosion,
weathering, efflorescence, discoloration, rotting, and
fatigue.
The focus on this paper will be items one and two.
CURRENT TREATMENT OF SERVICEABILITY
ISSUES IN U.S. CODES
A review of the three model building codes
3,29,35
in the United
States reveals a somewhat inconsistent and haphazard ap-
proach to serviceability issues. For instance, it is implied that
the codes exist strictly to protect life safety of the general
public. Yet, traditionally they have contained provisions for
deflection control of floor members while ignoring provisions
for other member types (columns, walls, mullions, etc.). No
mention is made of limits for wind drift, vibration, expansion
and contraction (expansion joint guidelines), or corrosion.
The authors work in professional committees and code
bodies, coupled with a review of recent surveys of the profes-
Lawrence G. Griffis is Senior Vice President and Director of
Structural Engineering for Walter P. Moore and Associates,
Houston, TX.
Serviceability Limit States Under Wind Load
LAWRENCE G. GRIFFIS
FIRST QUARTER / 1993 1
sion
36
seem to reveal a reluctance of engineers to codify
serviceability issues. This reluctance probably stems in part
on differences of opinion as to the purpose of building codes
(i.e., protection for life safety exclusively or establishment of
complete minimum design standards including strength and
serviceability), but also a genuine concern for restricting
design options, stifling creativity, and removing the all-
important concept of engineering judgment from the solu-
tion to the problem. There is also the belief, rightly so, that
too little hard data exists to justify rigid standards on most
serviceability issues.
It is important that engineers recognize these problems and
begin to focus on the solution of serviceability related design
issues. The reason for doing so is the large economic impact
that serviceability items are having on the operational costs
of buildings.
MEAN RECURRENCE INTERVAL WIND LOADS
FOR SERVICEABILITY DESIGN
The first step in establishing a serviceability design criterion
is to define the load under which it is to be checked. Wind
loading criteria for strength limit states in the United States
are normally based on a 50-year mean recurrence interval for
normal buildings and a 100-year mean recurrence interval for
critical structures. There seems to be a general consensus that
basing serviceability criteria on such a severe loading that
may occur only once, on the average, during the lifetime of
the structure is unrealistic and too stringent a standard to
apply. The average tenant occupancy in office buildings has
been defined as eight years.
26
It seems reasonable to base
serviceability criteria on a mean recurrence interval more in
this range of time because the consequences of exceeding a
serviceability limit state are usually not safety related. Various
researchers have suggested mean recurrence intervals of from
five to ten years for serviceability issues.
10,11,12,14,17,18,19,20,33,36
If no permanent damage results from exceeding the service-
ability limit, some researchers have also suggested selecting
serviceability criteria (such as floor deflection) on an annual
basis.
14
A wind load for a mean recurrence interval of 10 years is
recommended for checking the two wind serviceability limit
states defined herein (deformation and motion perception).
This corresponds to a 10 percent probability of being ex-
ceeded in any given year. While it has become standard
practice to base building accelerations under wind load on this
mean recurrence interval, drift criteria typically have been
formulated around the same mean recurrence interval (50
years or 100 years) as the strength limit state.
36
The proposed 10-year mean recurrence interval compares
to five years as proposed in ISO Standard 6897-1984, 10 years
as proposed by the National Building Code of Canada (1990),
20 years in the Australian Standard AS 1170.2-1989 and 0.1
years as proposed by the Japanese.
28
BUILDING DRIFTSTANDARD OF PRACTICE
Serviceability of buildings under wind loads has traditionally
been checked in the design office by evaluation of the lateral
frame deflection calculated on the basis of a statically applied
wind load obtained from the local building code. The magni-
tude of the wind load is usually the same as that used in
proportioning the frame for strength and typically is based on
a 50-year or 100-year mean recurrence interval load. Some-
times, an arbitrary wind load (i.e., 20 PSF above 100 ft, 0
(zero) PSF below 100 ft as has been used in New York City
on the design of some buildings
15
) is used in the serviceability
check. This serviceability check, for all but the tallest and
most slender of buildings (where wind tunnel studies are
utilized), has been used to prevent damage to collateral build-
ing materials, such as cladding and partitions, and also to
control the perception of building motion. None of the three
national building codes in the United States specify a limit to
lateral frame deflection under wind load. The degree of this
serviceability check is left to the judgment of the design
engineer. Lateral frame deflection is usually evaluated for the
building as a whole, where the applicable parameter is total
building drift, defined as the lateral frame deflection at the
top-most occupied floor divided by the height from grade to
the uppermost floor ( / H); and for each floor of the building,
where the applicable parameter is interstory drift, defined as
the lateral deflection of a floor relative to the one immediately
below it divided by the distance between floors ((
n

n1
) /
h). Typical values of this parameter (commonly called drift
index) used in this serviceability check are H / 100 to H / 600
for total building drift and h / 200 to h / 600 for interstory drift
depending on building type and materials used. The most
widely used values are 1 / 400 to 1 / 500.
36
Lateral frame
deflections have historically been based on a first order analysis.
DRIFTA REVISED DEFINITION
7
Drift Measurement Index (DMI)
If the goal in defining a drift limit is limited to only the control
of damage to collateral building elements, such as cladding
and partitions, and is separated from the problem of building
motion, then frame racking or shear distortion (strain) is the
logical parameter to evaluate.
Mathematically, if the local x, y displacements are known
at each corner of an element or panel, then the overall average
shear distortion for rectangular panel ABCD as shown in
Figure 1 may be termed the drift measurement index (DMI)
and defined as follows:
Drift measurement index (DMI) = average shear distortion
DMI = 0.5 [(X
A
X
C
) / H + (X
B
X
D
) / H + (Y
D
Y
C
) / L
+ (Y
B
Y
A
) / L]
DMI = 0.5 (D1 + D2 + D3 + D4)
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 2
where,
X
i
= vertical displacement of point i
Y
i
= lateral displacement of point i
D1 = (X
A
X
C
) / H, horizontal component of racking drift
D2 = (X
B
X
D
) / H, horizontal component of racking drift
D3 = (Y
D
Y
C
) / L, vertical component of racking drift
D4 = (Y
B
Y
A
) / L, vertical component of racking drift
It is to be noted that terms D1 and D2 are the horizontal
components of the shear distortion or frame racking and are
the familiar terms commonly referred to as interstory drift.
The terms D3 and D4 are the vertical components of the shear
distortion or frame racking caused by axial deformation of
adjacent columns.
If it can be accepted that the DMI is the true measure of
potential damage, then it becomes readily apparent that the
evaluation of interstory drift alone can be misleading in
obtaining a true picture of potential damage. Interstory drift
alone does not account for the vertical component of frame
racking in the rectangular panel that also contributes to the
potential damage, nor does it exclude rigid body rotation of
the rectangular panel which, in itself, does not contribute to
damage. It can be shown that evaluation of the commonly
used interstory drift can significantly underestimate the dam-
age potential in a combined shear wall/frame type building
where the vertical component of frame racking can be impor-
tant; and significantly overestimate the damage potential in a
shear wall or braced frame building where large rigid body
rotation of a story can occur due to axial shortening of
columns.
7
Consider for example, the eight-story building shown in
Figure 2. This frame represents a typical windframe that may
be found in any office building with 36-ft lease depths (build-
ing perimeter to center core) and a central core containing
elevator, stairs, etc. The frame shown consists of a combined
moment frame and X-braced frame. Figure 3 shows a plot
(exaggerated) of the deflected shape of the top level under
wind loads. Table 1 shows calculations for the traditional
story drift and the revised drift definition DMI. The signifi-
cant thing to note is that the potential damaging deformations,
as represented by the DMIs, are more severe in the external
bays (panels 1, 3) and much less severe in the internal bay
(panel 2) than predicted by the traditional story drift calcula-
tion. Most of the deformation in the center bay (panel 2) is
simply rigid body rotation that, by itself, is not damaging to
partitions.
Drift Measurement Zone (DMZ)
It is logical to identify the rectangular panel ABCD in Fig-
ure 1 as the zone in which the damage potential is to be
evaluated and define it the drift measurement zone (DMZ).
From a practical standpoint, these zones will typically repre-
sent column bays within a building and would be incorporated
as part of the building frame analysis.
Drift Damage Index (DDI)
Once the determination of the shear distortion or drift meas-
urement index (DMI) is made for different column bays or
drift measurement zones (DMZs), it must be compared to a
damage threshold value for the element being protected.
These damage threshold limits can be defined as the shear
distortion or racking that produces the maximum amount of
cracking or distress that can be accepted, on the average, once
Fig. 1. Drift measurement index (DMI). Figure 3
Figure 2
FIRST QUARTER / 1993 3
every 10 years. It is logical to define these damage threshold
shear distortions as the drift damage index (DDI). From the
standpoint of serviceability limit states it is necessary to
observe the following inequality:
drift measurement index dr iftdamageindex
DMI DDI
A significant body of information is available from racking
tests for different building materials that may be utilized to
define DDIs.
2
This is discussed further below in conjunction
with Figure 4.
Calculation of Building Frame Deflection
If drift measurement indices (DMIs) are to be effective in
controlling collateral building material damage, there must be
a consistency and accuracy in the method of calculation. A
recent survey
36
on drift clearly pointed out the problems that
exist in the structural engineering community on controlling
damage by excessive drift. There appears to be a wide vari-
ation in the methods of structural analysis performed to
calculate building frame deflection. Ideally, if DMIs are to be
an effective parameter in controlling damage caused by build-
ing deflection, then the structural analysis employed must
reasonably capture the significant response of the building
frame under load. As previously stated, it is suggested that the
wind load be defined by the 10-year mean recurrence interval
storm. The designer should recognize that the wind loads used
in the structural analysis are static equivalent wind loads
that are intended to estimate the peak load effect (mean plus
dynamic) caused by the vibratory nature of the building
motion. The structural analysis must then capture all signifi-
cant components of potential frame deflection as follows:
1. Flexural deformation of beams and columns.
2. Axial deformation of columns.
3. Shear deformation of beams and columns.
4. Beam-column joint deformation.
5. Effect of member joint size.
6. P- effect.
The behavioral knowledge of each of the above effects on
frame deflection is sufficiently understood to permit a reason-
ably accurate prediction of the contribution to the total re-
sponse. Computer programs and analytical models are now
within reach of most engineers to afford consideration of all
of the above effects.
Depending on the height, slenderness, and column bay
geometry, each of these effects can have a significant influ-
ence on building deflection. A recent study
8
on the sources of
elastic deformation for different height (10 to 50 stories) and
number-of-bay (5 to 13 bays) frames showed the following:
1. Axial deformations in columns can be very significant
for tall slender frames, amounting to 26 percent to 59
percent of the total deflection, depending on bay widths.
2. Shear deformations, as a percentage of the total frame
deflection, tend to increase with the number of bays and
also as the bay size (beam span) reduces. Shear defor-
mation can account for as much as 26 percent of the total
deflection. For slender tube structures (10- to 15-ft
bays and 40 to 50 stories tall) shear deformation can
contribute as much as flexural deformation to the total
building deflection. Shear deformations should never be
ignored in frame deflection if an accurate response pre-
diction is expected.
3. Beam-column joint deformations, particularly for steel
structures, constitute a significant portion of the total
deflection for all frames studied and should never by
ignored. As with shear deformations, there is a general
trend for deformations to increase as the number of bays
increases and the size of the bay decreases. Participation
Table 1.
Drift Comparison
D1 D2 D3 D4 Drift DMI DMI/Story Drift
Panel 1 0.00101 0.00104 0.000220 0.000215 0.00101 0.0012500 1.23
Panel 2 0.00104 0.00104 0.001030 0.001020 0.00101 0.0000186 0.02
Panel 3 0.00104 0.00101 0.000214 0.000209 0.00101 0.0012400 1.22
Fig. 4. Drift damage thresholdpartitions.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 4
as a percentage of the total varied from 16 percent to 41
percent.
4. The P- effect can easily increase total frame dis-
placement by 10 to 15 percent depending on frame
slenderness.
Errors in the determination of frame stiffness can also
affect proper design for strength. For example, the P- effect
is a function of frame stiffness; the magnitude of wind forces
in tall buildings is affected by building period; and the mag-
nitude of seismic forces is also affected by building period.
DAMAGE THRESHOLDS FOR
BUILDING MATERIALS
General guidelines to the behavior that might be expected
from different building elements and materials at various drift
indices may be obtained from a review of the literature.
2,13,31
A summary of behavior, taken from a recent study on ser-
viceability research needs
31
is shown in Table 2. Another
source of information may be found in seismic racking tests
of exterior cladding systems for buildings sometimes per-
formed during routine testing of mock-ups at testing labora-
tories. One of the most comprehensive studies of damage
intensity as a function of shear distortion can be found in
Reference 2 which contains a summary of over 700 racking
tests on various nonstructural partitions taken from more than
30 different sources. Partition types surveyed included tile
and hollow brick, concrete block, brick and veneer; walls
which consisted of gypsum wall board, plaster, and plywood.
Veneer walls are often referred to as drywall in engineering
practice. Damage intensity was defined on a scale from 0.0
to 1.0 with 0.1 to 0.3 defined as minor damage, 0.4 to 0.5
defined as moderate damage, 0.6 to 0.7 defined as substantial
damage, and 0.8 to 1.0 defined as major damage. A damage
intensity of 1.0 is defined as complete or intolerable. Figure 4
shows a plot of damage intensity versus shear distortion for
the partition groups discussed. If the upper limit of the minor
damage range is selected as the maximum acceptable dam-
age to occur in a 10-year design period, then the deflection
limit of 0.25 percent (1 / 400) is obtained for veneer or
drywall in Figure 4. This number correlates reasonably well
with the first damage threshold limit of
1

4
-in. displacement
for an eight foot tall test panel as described in Reference 13
for gypsum wallboard. The 0.3 damage intensity has been
used as the maximum acceptable shear distortion for the
various partition types in Table 3.
SERVICEABILITY LIMIT STATEDEFORMATION
(CURVATURE, DEFLECTION, DRIFT)
Once a wind load (mean recurrence interval) has been defined
for use in the serviceability check, the appropriate deforma-
tion to measure it has been defined (drift measurement index
(DMI)) and damage thresholds are determined from tests or
estimated, it remains only to establish an appropriate limit for
different building components. Table 3 is a compilation of
most common building elements with recommended defor-
mation limits. The building elements considered include roof,
exterior cladding, interior partitions, elevators, and cranes.
Most of the more common building cladding and partition
types are considered. Deformation types addressed include
deflection perpendicular to the plane of the building element
and shear deformation (racking) in the plane of the element.
Table 2.
Serviceability Problems at Various Deflection or Drift Indices
14,31
Deformation as a
Fraction of Span or
Height
Visibility of
Deformation Typical Behavior
1 / 1000 Not Visible Cracking of brickwork
1 / 500 Not Visible Cracking of partition walls
1 / 300 Visible General architectural damage
Cracking in reinforced walls
Cracking in secondary members
Damage to ceiling and flooring
Facade damage
Cladding leakage
Visual annoyance
1 / 200 1 / 300 Visible Improper drainage
1 / 100 1 / 200 Visible Damage to lightweight partitions,
windows, finishes
Impaired operation of removable
components such as doors,
windows, sliding partitions
FIRST QUARTER / 1993 5
Building Element
Supporting Structural
Element Deformation Type
Recommended
Limit Comments
Roof Membrane Roof
Metal Roof
Skylights
Purlin, Joist, Truss
Purlin, Joist, Truss
Purlin, Joist, Truss
Deflection Roof Plane
Deflection Roof Plane
Differential Support Deflection
L / 240
L / 150
L / 240
1

2
-in.

Note 1
Note 2
Exterior
Cladding
Brick Veneer Metal/Wood Stud
Horizontal Girts
Vertical Girts/Cols.
Wind Frame
Deflection Wall Plane
Deflection Wall Plane
Deflection Wall Plane
Shear Strain (DMI)
H / 600
L / 300
L / 600
H / 400
Note 3
Note 4
Note 4
Note 5
Concrete Masonry
Unreinforced
(Note 6)
Horizontal Girts
Vertical Girts/Cols.
Wind Frame, One-story
Wind Frame, Multi-story
Deflection Wall Plane
Deflection Wall Plane
Shear Strain (DMI)
Shear Strain (DMI)
L / 300
L / 600
H / 600
H / 400
Note 4
Note 4
Note 7
Note 8
Concrete Masonry
Reinforced
(Note 6)
Horizontal Girts
Vertical Girts/Cols.
Wind Frame, One-story
Wind Frame, Multi-story
Deflection Wall Plane
Deflection Wall Plane
Shear Strain (DMI)
Shear Strain (DMI)
L / 240
L / 240
H / 200
H / 400


Note 9
Note 10
Tilt-up
Concrete
Horizontal/Vertical Girts
Wind Frame
Deflection Wall Plane
Shear Strain (DMI)
L / 240
H / 200
Note 11
Note 12
Plaster,
Stucco
Metal/Wood Stud
Horizontal/Vertical Girts
Wind Frame
Deflection Wall Plane
Deflection Wall Plane
Shear Strain (DMI)
H / 600
L / 600
H / 400
Note 13
Note 13
Note 14
Architectural Precast
Concrete Panels,
Stone Clad Precast
Concrete Panels
Horizontal/Vertical Girts
Wind Frame
Deflection Wall Plane
Shear Strain (DMI)
L / 240
H / 400
Note 11
Note 15
Architectural Metal
Panel
Metal Stud,
Vertical/Horizontal Girts
Wind Frame
Deflection Wall Plane
Shear Strain (DMI)
L / 120
H / 100
Note 16
Note 17
Curtain Wall,
Window Wall
Mullions,
Horizontal/Vertical Girts
Wind Frame
Deflection Glass Plane
Shear Strain (DMI)
L / 175
H / 400
Note 18
Note 19
Interior
Partitions
Gypsum Drywall,
Plaster
Wind Frame Shear Strain (DMI) H / 400 Note 20
Concrete Masonry
Unreinforced
(Note 6)
Wind Frame Shear Strain (DMI) H / 667 Note 20
Concrete Masonry
Reinforced
(Note 6)
Wind Frame Shear Strain (DMI) H / 400 Note 10, 20
Tile, Hollow Clay
Brick
Wind Frame Shear Strain (DMI) H / 2000 Note 20
Brick Wind Frame Shear Strain (DMI) H / 1250 Note 20
Table 3.
Wind Serviceability Limit State Deformation
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 6
Table 3, contd
Wind Serviceability Limit State Deformation
Building Element
Supporting Structural
Element Deformation Type
Recommended
Limit Comments
Elevators Wind Frame Shear Strain (DMI) H / 400 Note 21
Cranes Cab Operated Wind Frame Shear Strain (DMI) H / 240 2-in. Note 22
Pendant Operated Wind Frame Shear Strain (DMI) H / 100 Note 23
Notes to Table C.
H = story height L = span length of supporting member DMI = drift measurement index
1. Metal roofs include standing seam and thru fastener type roofs.
12
2. Deflection limit shown is relative support movement measured perpendicular to a line drawn between skylight support points. Racking movements in the plane
of the glass should be limited to
1
4-in. for gasketed mullions and
1
8-in. for flush (butt) glazing.
12
3. Deflection limits recommended by the Brick Institute of America
34
are L/600 L/720.
4. L/600 is recommended for the case when predominant flexural stress in masonry is perpendicular to bed joint. L/300 may be used for the case when
predominant flexural stress in masonry is parallel to bed joint.
5. H/400 limit applies if brick is supported on relief angles at each floor with
3
8-in. soft joint and
3
8-in. control joints are used in each column bay.
6. Reinforced concrete masonry implies vertical reinforcing bars in grouted cells and/or horizontal reinforcing bars in bond beams.
7. Assumes only windframe designed to carry lateral loads and flexible connections used between wall and parallel windframe. H/600 limit also protects wall
perpendicular to plane of windframe from excessive flexural cracking. A horizontal crack control joint at base of wall is recommended. Limit crack width
under wind load to
1
16-in. if no joint is used and
1
8-in. if control joint is used.
12
8. Assumes only windframe designed to carry lateral loads and flexible connections used between wall and parallel windframe. H/400 applies only if in-fill walls
have
3
8-in. soft joints against structural frame.
9. Assumes only windframe designed to carry lateral loads and flexible connections used between wall and parallel windframe. Stricter limit should be considered
if required to protect other building elements. If walls designed as shear walls, then design DMI should be based on damage control of other building elements.
H/200 limit also protects wall perpendicular to plane of windframe from excessive flexural cracking. If a horizontal control joint at base of wall is used, then
limit may be changed to H/100.
12
10. H/400 limit applies to reinforced masonry walls designed as shear walls unless stricter limit is required to protect other more critical building elements. Reinforced
masonry walls infilled hard against structural windframe should not be used without assessing their stiffness in a compatibility analysis with windframe, unless
isolation joints are provided between wall and building frame.
11. In cases where wall support is indeterminate, differential support deflection should be considered in design of wall panel.
12. Assumes only windframe designed to carry lateral loads and flexible connections are used between wall and parallel windframe. Stricter limit should be
considered if required to protect other building elements. If panels designed as shear walls then H/400 is recommended limit with minimum
3
4-in. panel joints.
13. Control joints are recommended to limit cracking from shrinkage, thermal, and building movement.
14. H/400 limit applies if wall is panelized with
3
8-in. control joints and relief joints are used between floors and at each column bay. If plaster applied to unreinforced
masonry, then limits should be same as masonry.
15. H/400 applies if panel connection to frame is determinate, flexible connections are used between panel and parallel windframe and minimum
3
4-in. panel joints
are used. Panels with indeterminate support to frame should be designed for differential support movement.
16. Consult metal panel manufacturer for possible stricter requirements.
17. L/100 limit applies for metal panel only. Other building components may warrant stricter limit.
18. L/175 recommended by American Architectural Manufacturers Association.
27
Recommended limit changes to L/360 when a plastered surface or dry wall
subjected to bending is affected. At roof parapet or other overhangs recommended limit is 2L/175 except that the deflection of a member overhanging an
anchor joint with sealed joint (such as for roof flashing, parapet cover, soffit) shall be limited to no more than one half the sealant joint depth between the
framing member and fixed building element.
19. H/400 limit is to protect connections to building frame and also sealants between panels. More liberal limits may be applicable for custom designed
curtain/window walls where racking can be accounted for in design and where wall will be tested in a labortory mock-up. Consult manufacturer for racking
limits of off-the-shelf systems.
20. Recommended limits shown assume partition is constructed hard against structural frame. More liberal limits may be appropriate if isolation (soft) joints are
designed between partition edge and structural frame. Design of structural frame for DMI limits stricter than H/600 is probably not practical or cost effective.
21. In addition to the static deflection limit shown, proper elevator performance requires consideration of building dynamic behavior. Design of elevator systems
(guide rails, cables, sheaves) will require knowledge of predominant building frequencies and amplitude of dynamic motion. This information should be furnished
on the drawings or in the specifications.
22. Limit shown applies to wind loads or crane forces, either lateral or longitudinal to crane runway. Deflection limit specified is to be measured at the elevation of
crane runways.
12
23. Buildings designed to H/100 limit will exhibit observable movements during crane operation. Stricter limits may be appropriate to control this and/or to protect
other building components.
12
FIRST QUARTER / 1993 7
Notes are included at the end of the table to explain or clarify
a recommendation.
It should be pointed out that the recommended limits
shown are guidelines based on past successful performance.
The degree of distress in any of the building elements (clad-
ding, partitions) under the action of wind loads is highly
dependent upon the nature and design of the attachments or
joints to the building frame. If specific attention is paid to this
aspect then oftentimes any reasonable deformation can be
accommodated without damage. Indeed, it may be more
prudent and cost effective to detail joints to accommodate a
higher deflection than to design a higher level of stiffness into
the building wind frame.
SERVICEABILITY LIMIT STATE
MOTION PERCEPTION
Motion Perception ParameterAcceleration
28,30
Perception to building motion under the action of wind may
be described by various physical quantities including maxi-
mum values of velocity, acceleration, and rate of change of
acceleration, sometimes called jerk. Since wind induced mo-
tion of tall buildings is composed of sinusoids having a nearly
constant frequency f but varying phase, each quantity is
related by the constant 2f where f is the frequency of motion
(V = 2 f )D; A = (2 f )
2
D; J = (2 f )
3
D where D, V, A, and
J are maximum displacement, velocity, acceleration, and jerk
respectively). Human response to motion in buildings is a
complex phenomenon involving many psychological and
physiological factors. It is believed that human beings are not
directly sensitive to velocity if isolated from visual effects
because, once in motion at any constant velocity, no forces
operate upon the body to keep it in such motion. Acceleration,
on the other hand, requires a force to act which stimulates
various body organs and senses. Some researchers believe
that the human body can adapt to a constant force acting upon
it whereas with changing acceleration (jerk) a continuously
changing bodily adjustment is required. This changing accel-
eration may be an important component of motion perception
in tall buildings. It appears that acceleration has become the
standard for evaluation of motion perception in buildings
because it is the best compromise of the various parameters.
It also is readily measurable in the field with available equip-
ment and has become a standard for comparison and estab-
lishment of motion perception guidelines among various re-
searchers around the world.
Factors Affecting Human Response to Building Motion
25
Perception and tolerance thresholds of acceleration as a meas-
ure of building motion are known to depend on various factors
as described below. These factors have been determined from
motion simulators that have attempted to model the action of
buildings subjected to wind loads.
1. Frequency or Period of Building. Field tests have
shown that perception and tolerance to acceleration
tend to increase as the building period increases (fre-
quency decreases) within the range of frequency com-
monly occurring in tall buildings.
2. Sex. The general trend of response between men and
women is the same although women are slightly more
sensitive than men.
3. Age. The sensitivity of humans to motion is an inverse
function of age, with children being more sensitive
than adults.
4. Body Posture. The sensitivity of humans to motion is
proportional to the distance of the persons head from
the floor; the higher the persons head, the greater the
sensitivity. Thus, a persons perception increases as he
goes from sitting on the floor, to sitting in a chair, to
standing. However, since freedom of the head may be
important to motion sensitivity, a person sitting in a
chair may be more sensitive than a standing person
because of the body hitting the back of the chair.
5. Body Orientation. Humans tend to be more sensitive to
fore-and-aft motion than to side-to-side motion be-
cause the head can move more freely in the fore-and-aft
direction.
6. Expectancy of Motion. Perception threshold decreases
if a person has prior knowledge that motion will occur.
Threshold acceleration for the case of no knowledge is
approximately twice that for the case of prior knowledge.
7. Body Movement. Perception thresholds are higher for
walking subjects than standing subjects, particularly if
the subject has prior knowledge that the motion will
occur. The perception threshold is more than twice as
much between the walking and standing case if there
is prior knowledge of the event, but only slightly
greater if there is no knowledge of the event.
8. Visual Cues. Visual cues play an important part in
confirming a persons perception to motion. The eyes
can perceive the motion of objects in a building such
as hanging lights, blinds, and furniture. People are also
very sensitive to rotation of the building relative to
fixed landmarks outside.
9. Acoustic Cues. Buildings make sounds as a result of
swaying from rubbing of contact surfaces in frame
joints, cladding, partitions, and other building ele-
ments. These sounds and the sound of the wind whis-
tling outside or through the building are known to focus
attention on building motion even before subjects are
able to perceive the motion, and thus lower their per-
ception threshold.
10. Type of Motion. Under the influences of dynamic wind
loads, occupants of tall buildings can be subjected to
translational acceleration in the x and y direction and
torsional acceleration as a result of building oscillation
in the along-wind, across-wind, and torsional direc-
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 8
tions, respectively. While all three components con-
tribute to the response, angular motion appears to be
more noticeable to occupants, probably caused by an
increased awareness of the motion from the aforemen-
tioned visual cues. Also, torsional motions are often
perceived by a visual-vestibular mechanism at motion
thresholds which are an order of magnitude smaller
than those for lateral translatory motion.
24
Root-Mean-Square (RMS) Versus Peak Acceleration
A review of the literature on the subject of motion perception
as measured by acceleration shows a difference in the pres-
entation of the results. Some researchers report maximum or
peak acceleration and some report root-mean-square or RMS
accelerations. This dual definition has extended into estab-
lishing standards for motion perception.
Most of the research conducted on motion perception has
been with motion simulators subjected to sinusoidal motion
with varying frequency and amplitude. In these tests it has
been common to report the results in maximum or peak
acceleration since that was the quantity directly measured. It
should be pointed out that for sinusoidal acceleration, the
peak is equal to 2 times the RMS value. It appears that wind
tunnel research has tended to report peak acceleration or both
peak and RMS in order to correlate the wind tunnel studies
with these motion simulation tests. Many researchers believe
that, when the vibration persists for an extended period of
time (10 to 20 minutes) as is common with windstorms having
a 10-year mean recurrence interval, that RMS acceleration is
a better indicator of objectionable motion in the minds of
building occupants than isolated peak accelerations that may
be dampened out within a few cycles.
17,18,22,33
Also, the RMS
statistic is easier to deal with during the process of temporal
and spatial averaging because the 20-minute averaging period
for a storm represents a time interval over which the mean
velocity fluctuates very little. The relationship between peak
and RMS accelerations in tall buildings subjected to the
dynamic action of wind loads has been defined by the peak
factor which varies with building frequency, but which is
oftentimes taken as 3.5. Correlation between peak and RMS
accelerations in tall building motion may be made using this
peak factor.
Relationship Between Building Drift and
Motion Perception
Engineers of tall buildings have long recognized the need for
controlling annoying vibrations to protect the psychological
well being of the occupants. Prior to the advent of wind tunnel
studies this need was addressed using rule-of-thumb drift
ratios of approximately 1/400 to 1/600 and code specified
loads. Recent research,
22
based on measurement of wind
forces in the wind tunnel, has clearly shown that adherence
to commonly accepted lateral drift criteria, per se, does not
explicitly ensure a satisfactory performance with regard to
motion perception. The results of one such study
22
are plotted
in Figure 5 for two square buildings having height/width
ratios of 6/1 and 8/1 where each is designed to varying drift
ratios. Plots are shown of combined transitional and torsional
acceleration as a function of design drift ratio. At drift ratios
of 1/400 and 1/500 neither building conforms to acceptable
standards for acceleration limits. The reason that drift ratios
by themselves do not adequately control motion perception
is because they only address stiffness and do not recognize
the important contribution of mass and damping, which to-
gether with stiffness, are the predominant parameters affect-
ing acceleration in tall buildings. This is discussed further
later in the paper.
Human Response to Acceleration
Considerable research in the last 20 years has been conducted
on the subject of determining perception threshold values for
acceleration caused by building motion.
9,25,28
Much of this
work has also attempted to formulate design guidelines for
tolerance thresholds to be used in the design of tall and slender
buildings.
Some of the earliest attempts to quantify the problem were
performed by Chang
5,6
who proposed peak acceleration limits
for different comfort levels that were extrapolated from data
in the aircraft industry. Changs proposed limits, plotted in
Figure 6 as a function of building period, are stated as follows:
Peak Acceleration Comfort Limit
<0.5% g Not Perceptible
0.5% to 1.5% g Threshold of Perceptibility
1.5% to 5.0% g Annoying
5% to 15.0% g Very Annoying
>15% g Intolerable
Additional data has been reported by researchers who
utilized motion simulators to define perception levels.
9,28
A
summary of this work is shown in Figures 7 and 8 showing
plots of perception thresholds for both peak and RMS accel-
eration as a function of building period.
Perhaps the most comprehensive studies of the problem
have been performed in Japan
28
for a wide range of variables.
Fig. 5. RMS acceleration vs drift index.
FIRST QUARTER / 1993 9
This work is summarized in Figure 8 where peak acceleration
is plotted as a function of building period. Each curve and
zone between curves is identified in the figure. The discussion
below is keyed to the letters and numbers in the figure and is
taken from Reference 28:
1. Zone A, below Curve 1, identifies peak acceleration
less than about 0.5 percent g. In this zone, a human
cannot perceive motion at all. No evidence of motion
exists except for possible rubbing of building compo-
nent surfaces in contact. Curve 1 defines the limit of
perception threshold for an average population.
2. Curve 2 (0.5 percent g) defines the point where some
building objects (furniture, hanging lights, water) be-
gin to move.
3. Curve 3 separates zones between very normal walk-
ing and nearly normal walking.
4. Zone B (between 0.5 percent g and 1.0 percent g)
identifies a zone where some people can perceive
motion. Some building fixtures and objects will begin
to move slightly, but these movements are generally
not observable except to a person who looks directly
at them.
5. Curve 4 (1 percent g) separates the zones where people
can be affected by working at a desk.
6. Curve 5 defines the threshold where people can start to
become subjected to motion sickness when exposed to
this level of motion for extended periods.
7. Zone C (between 1.0 percent g and 2.5 percent g) is
where most people are able to perceive motion and
become affected by desk work. Generally, in this zone,
people can be subjected to motion sickness if exposed
for extended periods but can walk without hindrance.
8. Curve 6 defines the limit between normal and hindered
walking.
9. Zone D (between 2.5 percent g and 4.0 percent g)
defines the acceleration range where desk work be-
comes difficult and at times impossible. Most people
can walk and go up and down stairs without too much
difficulty.
10. Curve 7 (3.5 percent g) defines the point where working
at a desk is difficult.
11. Curve 8 (4.0 percent g) defines the acceleration where
furniture and fixtures begin to make sounds, which
may evoke a strong concern or alarm among some
people.
12. In Zone E people strongly perceive motion and stand-
ing people lose their balance and find it hard to walk
naturally.
13. Curve 9 marks the point where people are unable to
walk.
14. Curve 10 defines the maximum tolerance for motion.
15. In Zones F and G (above 5.0 percent g) most people
cannot tolerate the motion and are unable to walk.
These zones are considered to be at the limit of walking
ability.
16. In Zone H people cannot walk. Motion is intolerable.
Design of Tall Buildings for Acceleration
The design of most tall buildings is controlled by lateral
deflection and most often by perception to motion. Indeed,
this characteristic is often proposed as one definition of a
tall building.
While the problem of designing for motion perception in
tall buildings is usually solved by conducting a scale model
Fig. 7. Perception thresholdRMS acceleration. Fig. 8. Perception thresholdspeak acceleration.
Fig. 6. Tolerance thresholds proposed by Chang.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 10
force-balance or aeroelastic test in the wind tunnel, certain
criteria have been established to aid the designer. Empirical
expressions now exist
21,22,32
that allow approximate evalu-
ation of the susceptibility of a building to excessive motion.
This can be very helpful in the early design stages particularly
where geometry, site orientation, or floor plan are not yet
fixed.
The following simple expressions
22
for along-wind (drag),
across-wind (lift), and torsional RMS acceleration have been
derived for square, symmetric (coincident centers of mass,
rigidity, and geometry), tall buildings in an urban environ-
ment:
Along-wind:
A
D
(Z) = C
D
(Z)
U
H
2.74
K
D
0.37

0.5
M
D
0.63
(1)
Across-wind:
A
L
(Z) = C
L
(Z)
U
H
3.54
K
L
0.77

0.5
M
L
0.23
(2)
Torsional:
A

(Z) = C

(Z)
U
H
1.88
K

0.06

0.5
M

1.06

,

N

B
U
H
0.25 (3a)
A

(Z) = C

(Z)
U
H
2.76
K

0.38

0.5
M

0.62

,

N

B
U
H
> 0.25 (3b)
The proportionality constants C
D
(Z), C
L
(Z), and C

(Z) are
defined as follows:
C
D
(Z) = 0.0116 B
0.26
Z
C
L
(Z) = 0.0263 B
0.54
Z
C

(Z) = 0.00341 B
2.12
Z,
N

B
U
H
0.25
C

(Z) = 0.00510 B
1.24
Z,
N

B
U
H
> 0.25
The definition of terms in the above expressions are listed
below:
A
D
(Z), A
L
(Z), A

(Z) = along-wind, across-wind, and


torsional RMS acceleration at
height Z (meters/sec
2
, radians/sec
2
)
U
H
= mean hourly wind speed at the top of the building
(meters/sec.)
H = building height (meters)
B = plan dimension of square building (meters)
M = generalized mass of the building (kilogram)
=

m
i
i = 1
n

i
2
m
i
is mass of floor i and
i
is modal
coordinate at floor i, normalized so that = 1 at
(Z) = H
N = frequency (hertz)
K = generalized stiffness (newton/meters)
= (2N)
2
M
= damping ratio
For rectangular buildings, B may be taken as the square
root of the plan area. The resultant RMS acceleration at the
corner of the building, A
R
, is calculated as follows:
A
R
= (A
D
2
+ A
L
2
+ (B / 2 A

)
2
)
0.5
(4)
These expressions were used in a parametric study of a
150-ft square building having slenderness ratios (H / B) of
five through ten (building heights varying from 754 feet to
1,495 feet). The buildings were subjected to basic wind speed
of 70 mph in an Exposure B (suburban) environment as
defined in ASCE 7-88. The buildings were assumed to be
all-steel with steel weights typical of tall buildings of these
heights, varying from 25 psf to 44 psf. Building densities were
assumed to vary from 7.77 pcf to 9.23 pcf, typical for office
buildings having lightweight concrete metal deck floors and
curtain wall cladding. Translational building periods were
calculated using the well-known Rayleigh formula,
35
which
for uniform prismatic buildings with a linear deflected shape
can be approximated by the following expression:
T = 0.904H

D
R
pR

0.5
(5)
In this expression, T is the building period in seconds, H is
the building height (feet), is the density (PCF), D
R
is the
design drift ratio ( / H), p is the equivalent uniform pressure
(PSF) and R is the aspect ratio H / B. Torsional periods were
taken as 85 percent of the translational periods. For this study,
the drift ratio under design wind load as defined by ASCE
7-88 is set at 1/400 or 0.0025. This practice is typical of the
procedure used in many building designs.
Along-wind, across-wind, and torsional RMS accelera-
tions were calculated at the building top corner using 10-year
mean recurrence interval wind loads. Complete building data
is shown in Table 4 and the accelerations are plotted in
Figure 9. Also shown in Figure 9 is the design limit as defined
later in this paper. The results clearly show that controlling
drift limits to the traditional design value of 0.0025 does not
ensure satisfactory performance from the standpoint of mo-
tion perception. In examining Figure 9, it is interesting to note
that for the common aspect ratios of 5-6, torsional accelera-
tion is comparable to across-wind acceleration and both are
significantly larger than the along-wind acceleration.
Generally, for most tall buildings without eccentric mass
or stiffness, the across-wind response will predominate if
(WD)
0.5
/ H < 0.33 where W and D are the across-wind and
FIRST QUARTER / 1993 11
along-wind plan dimensions respectively and H is the build-
ing height.
32
In examining the across-wind proportionality, which often-
times is the predominant response, it is possible to make the
following deductions:
1. If stiffness is added without a change in mass, accelera-
tion will be reduced in proportion to 1 / N
1.54
, which is
proportional to 1 / K
0.77
, where K is the stiffness.
2. If mass is added throughout the building without chang-
ing the stiffness, acceleration will be reduced in propor-
tion to 1 / M
0.23
.
3. If mass is added with a proportionate increase in stiffness
Table 4.
Parametric Study
RMS Acceleration
150-ft. Square Building
H Ft. H / B
T
L
, T
D
(SEC)
T


(SEC)
STL. WT.
(PSF)

(PCF)
U
H
(MPH)
A
D
(Milli-g)
A
L
(Milli-g)
B / 2 A

(Milli-g)
A
R
(Milli-g)
754 5 6.85 5.82 25 7.77 58.6 2.96 4.60 4.83 7.30
897 6 7.31 6.21 29 8.08 63.3 3.69 6.43 6.04 9.57
1053 7 7.77 6.60 33 8.38 68.0 4.53 8.78 6.82 12.00
1196 8 8.19 6.96 37 8.69 71.8 5.26 11.10 7.22 14.25
1352 9 8.61 7.32 41 9.00 72.0 5.32 11.73 6.98 14.65
1495 10 8.97 7.62 44 9.23 72.0 5.35 12.18 6.78 14.93
NOTE: RMS accelerations are calculated using Equations 1, 2, and 3.
Table 5.
Traditional Motion Perception (Acceleration) Guidelines (Note 1)
10-year Mean Recurrence Interval
Occupancy
Type
Peak
Acceleration
(Milli-g)
Root-mean-square (RMS)
Acceleration (Milli-g)
1 T < < 4
0.25 < < f 1.0
( (g
p
4.0) )
4 T < < 10
0.1 < < f 0.25
( (g
p
3.75) )
T 10
f 0.1
( (g
p
3.5) )
Commercial
Residential
1527
Target 21
1020
Target 15
3.756.75
Target 5.25
2.505.00
Target 3.75
4.007.20
Target 5.60
2.675.33
Target 4.00
4.297.71
Target 6.00
2.865.71
Target 4.29
Notation:
T = period (seconds)
f = frequency (hertz)
gp = peak factor
NOTE:
1. RMS and peak accelerations listed in this table are the traditional unofficial standard applied in U.S. practice
based on the authors experience.
Fig. 9. Parametric study150 -ft sq bldg.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 12
so that N does not change, then the acceleration will be
reduced in proportion 1 / M or 1 / K.
4. If additional damping is added, then the acceleration will
be reduced in proportion to 1 /
0.5
.
It should be pointed out that torsional response can be
important even for symmetrical buildings with uniform stiff-
ness. This is because a torsional wind loading can occur from
unbalance in the instantaneous pressure distribution on the
building surface.
Oftentimes, in very slender buildings, it is not possible to
obtain satisfactory performance, given building geometry and
site constraints, by adding stiffness and/or mass alone. The
options available to the engineer in such a case involve adding
additional artificial damping and/or designing mass or pen-
dulum dampers to counteract the sway.
16
Standards of Motion Perception
Numerous high-rise buildings have been designed and are
performing successfully all over the world. Many have been
designed according to an unofficial standard observed in
the authors practice as defined in Table 5. Both peak accel-
eration and RMS accelerations are used, their relationship
generally defined by the use of a peak factor, g
p
, approxi-
mately 3.54.0. The true peak factor for a building which
relates the RMS loading or response to the peak, can be
determined in a wind tunnel aeroelastic model study.
32
Target
peak accelerations of 21 milli-gs and 15 milli-gs are often
used for commercial and residential buildings respectively.
Corresponding RMS values are ratioed accordingly using the
appropriate peak factor. A stricter standard is often applied to
residential buildings for the following reasons:
4
1. Residential buildings are occupied for more hours of the
day and week and are therefore more likely to experience
the design storm event.
2. People are less sensitive to motion when at work than
when in the home at leisure.
3. People are more tolerant of their work environment than
of their home environment.
4. Occupancy turnover rates are higher in office buildings
than in residential buildings.
5. Office buildings are more easily evacuated in the event
of a peak storm event.
The apparent shortcoming in the standard defined by
Table 5 is the fact that the tolerance levels are not related to
building frequency. Research has clearly shown a relationship
between acceptable acceleration levels and building fre-
quency. Generally higher acceleration levels can be tolerated
for lower frequencies (see Curves 1, 4 and 5 in Figure 7 and
Curves 5 and 6 in Figure 8).
The International Organization of Standardization has es-
tablished a design standard for occupant comfort in fixed
structures subjected to low frequency horizontal motion
ISO Standard 6897-1984.
17
This standard is based on a five-
year mean recurrence interval and seems to agree quite well
with the experimental work described in Figures 7 and 8. The
mean threshold curve from this standard is plotted for com-
parison to the research of Reference 9 in Figure 7. The ISO
Standard 6897 design curves are plotted in Figure 10. The
interesting feature of the ISO approach is that acceleration
limits increase as the building period increases and therefore
it represents a better correlation to available research. The
acceleration limits defined by the General Purpose Build-
Fig. 10. RMS accelerationISO 68971984
5-yr return period.
Fig. 12. Design standardpeak acceleration
10-yr return period.
Fig. 11. Design standardRMS acceleration
10 -yr return period.
FIRST QUARTER / 1993 13
ings curve of Figure 10 agree very well with U.S. practice
for commercial buildings if an upward adjustment of approxi-
mately 10 percent is used to account for the difference in mean
recurrence intervals for U.S. practice (10-year versus the ISO
five-year mean recurrence interval). The 10 percent adjust-
ment seems reasonable in light of the authors experience in
wind engineering studies performed on office buildings.
The authors observation and experience with U.S. prac-
tice, combined with a study of the available research pre-
viously described and also the ISO Standard 6897-1984, form
the basis of a proposed new standard defined in Figures 11
and 12. Design limits are proposed for both peak and RMS
acceleration using a 10-year mean recurrence interval wind
as customarily used in U.S. practice. The logic used in the
formulation of these curves is described below:
1. Design curves are established for residential buildings
and for commercial buildings. Residential buildings de-
mand a separate stricter standard for the reasons pre-
viously stated. Target values are given for each building
type centered between an upper and lower bound. The
upper bound values are 12.5 percent above and the lower
bound values 12.5 percent below the target values. The
concept of a design range seems reasonable considering
the limited available research and the uncertainty in the
present state-of-the-art.
2. The ISO 6897 curve for mean threshold acceleration
(middle curve of Figure 10) is taken as a lower bound
for the residential building curves shown in Figure 11.
3. The target and the upper bound values are established
considering the design range defined in Item 1.
4. The commercial building target curve is defined by
using the ISO Standard General Purpose Building
curve of Figure 10, increased by 10 percent to reflect the
change in mean recurrence intervals. The upper and
lower bounds are defined 12.5 percent above and below
the target curve respectively.
5. The peak acceleration curves defined in Figure 12 are
based on the corresponding RMS acceleration curves of
Figure 11 multiplied by a peak factor as defined in Table 5.
Additional research and experience will be required to
confirm the validity of this proposed new standard. The
acceleration levels relate reasonably well (slightly higher)
with the successful experience of Table 5 and the new stand-
ard has the advantage of frequency dependency that seems to
be confirmed by research.
CONCLUSIONS
This paper has focused on two serviceability limit states for
buildings (particularly tall and/or slender buildings); namely,
deformation (deflection, curvature, and drift) and motion
perception as measured by acceleration.
The conclusions reached in this paper are summarized
below:
1. The current practice of using 50-year or 100-year mean
recurrence interval wind loads to evaluate building
drift with currently accepted drift limits is overly con-
servative. Wind drift and acceleration are proposed to
be based on a mean recurrence interval of 10 years.
2. A revised definition of building drift is proposed to
better reflect the potential for damage to building ele-
ments. The new definition, termed herein as the drift
measurement index (DMI) is a mathematical formula-
tion of shear deformation or racking that occurs in a
building element. It includes the vertical component of
racking and filters out the effect of rigid body rotation,
both of which are shortcomings in the present defini-
tion of building drift. The term given to the rectangular
panel forming the zone over which shear deformation
is to be measured is drift measurement zone (DMZ).
The threshold damage distortion that represents the
limit of shear deformation that causes distress is termed
the drift damage index (DDI). The drift limit state may
then be stated as DMI DDI under 10-year wind loads.
3. If rational drift limits are to be effective, the calculation
of building drift must capture all significant compo-
nents of frame deflection including flexural deforma-
tion of beams and columns, axial deformation of col-
umns, shear deformation of beams and columns,
beam-column joint deformation (panel zone deforma-
tion), effect of member joint size, and the P- effect.
4. A review of available racking distortion data for differ-
ent partition types is made. Based on this information,
and past successful experience, guidelines (Table 3)
are proposed for different building elements (roofs,
cladding, partitions, elevators, and cranes) subjected to
10-year wind loads.
5. Factors affecting human response to building motion
are reviewed and include building frequency, sex, age,
body posture, body orientation, expectancy of motion,
and body movement of the occupants; visual cues,
acoustic cues, and type of motion.
6. Acceleration appears to be the best indicator of build-
ing motion at present.
7. Both RMS (root-mean-square) and peak acceleration
values are commonly used to represent building mo-
tion. There appears to be a difference of opinion among
engineers and researchers as to the relative importance
and merits of each. This issue should be resolved to
avoid confusion in the development of design
standards.
8. Contrary to early attempts by engineers to control
annoying lateral vibrations in buildings, building stiff-
ness, represented by drift ratios, by itself is not a good
indicator of occupant susceptibility to building motion
(Figure 5). Perception of building motion is influenced
by available damping and also building mass as well
as building stiffness.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 14
9. Research seems to indicate that human perception to
acceleration begins at about 0.5 percent (peak accel-
eration) and appears to increase as the building period
increases (Figure 8).
10. Human tolerance to acceleration tends to increase with
building period above about three to four seconds
(Figure 8).
11. Current practice in tall building design has targeted
design values for acceleration at 21 milli-gs peak
acceleration (6 milli-gs RMS) for office buildings and
15 milli-gs peak acceleration (4.3 milli-gs RMS) for
residential buildings. These limits do not recognize the
apparent trend for the dependence of acceleration lim-
its on building period.
12. Humans appear to be particularly sensitive to torsional
acceleration and so this component should be mini-
mized in the assignment of building mass and stiffness
as much as possible during the design stage.
13. The factors affecting building acceleration are best
evaluated in a wind tunnel study. Acceleration involves
the complex inter-relationship of the variables of mass,
stiffness, and damping and also the influence of build-
ing orientation on the site and the surrounding wind
environment.
14. Most tall slender building motion is controlled by
across-wind effects (vortex shedding). Generally
speaking, this component of acceleration is propor-
tional to the wind velocity to a power of about 3.5 and
the period of the building to a power of about 1.5; and
inversely proportional to mass and the square root of
damping.
15. Proposed standards for building perception (commer-
cial and residential buildings) are shown in Figures 11
and 12 which show acceleration limits increasing with
building period. This seems to follow the results of past
research and is an improvement over current standards.
16. The current approach to serviceability design is incon-
sistent in U.S. building codes and seems to reflect a
general reluctance by practicing engineers to codify
serviceability standards.
17. Structural engineers must begin to address service-
ability issues in design and establish rational standards
because of the increasing economic impact service-
ability issues are having on construction.
REFERENCES
1. American Institute of Steel Construction, Load and Resis-
tance Factor Design, AISC, Chicago, 1986.
2. Algan, B., Drift and Damage Considerations in Earth-
quake Resistant Design of Reinforced Concrete Build-
ings, Ph.D. Dissertation, Department of Civil Engineer-
ing, University of Illinois at Urbana, 1982.
3. BOCA National Building Code/1990, Building Officials
and Code Administrators International, Inc., Eleventh
Edition, 1990.
4. Cermak J., Boggs, D., Paper to be published.
5. Chang, F., Wind and Movement in Tall Buildings, Civil
Engineering Magazine, ASCE, August 1967.
6. Chang, F. K., Human Response to Motions in Tall Build-
ings, Paper presented at ASCE National Environmental
Engineering Meeting, Houston, Texas, October 1622.
7. Charney, F. A., Wind Drift Serviceability Limit State
Design of Multi-story Buildings, Journal of Wind Engi-
neering and Industrial Aerodynamics, 36, 1990.
8. Charney, F. A., Sources of Elastic Deformation in Later-
ally Loaded Steel Frame and Tube Structures, Council
on Tall Buildings and Urban Habitat, Fourth World Con-
gress, Tall Buildings: 2000 and Beyond, Hong Kong,
November 59, 1990.
9. Chen, P. W., Robertson, L. E., Human Perception
Thresholds of Horizontal Motion, Journal of Structural
Engineering, ASCE, Vol. 98, No. ST8, August 1972.
10. Davenport, A. G., Tall BuildingsAn Anatomy of Wind
Risks, Construction in South Africa, December 1975.
11. Ellingwood, B., Serviceability Guidelines for Steel
Structures, Engineering Journal, AISC, Volume 26, No.
1, First Quarter, 1989.
12. Fisher, J. M., West, M. A., Serviceability Design Consid-
erations for Low Rise Buildings, Steel Design Guide
Series No. 3, AISC, 1990.
13. Freeman, S. A., Racking Tests of High Rise Building
Partitions, Journal of Structural Engineering, ASCE,
Volume 103, No. ST8, August 1977.
14. Galambos, T. V., Ellingwood, B., Serviceability Limit
State: Deflection, Journal of Structural Engineering,
ASCE, Volume 12, No. 1, January 1986.
15. Gaylord, Edwin H. Jr., Gaylord, Charles N., Structural
Engineering Handbook, McGraw Hill Book Company,
New York, 1979.
16. Grossman, J. S., Slender Concrete StructuresThe New
Edge, ACI Structural Journal, Volume 87, No. 1, Janu-
ary/February 1990.
17. Guidelines for the Evaluation of the Response of Occu-
pants of Fixed Structures to Low Frequency Horizontal
Motion (0.063 to 1 Hz.), ISO Standard 6897-1984, In-
ternational Organization of Standardization, 1984.
18. Hansen, R. J., Reed, J. W., Vanmarcke, E. H., Human
Response to Wind-Induced Motion of Buildings, Jour-
nal of Structural Engineering, ASCE, Volume 99, No.
ST7, July 1973.
19. Irwin, A. W., Human Response to Dynamic Motion of
Structures, Structural Engineer, Volume 56A, No. 9,
September 1978.
20. Irwin, A. W., Motion in Tall Buildings, Paper presented
at Second Century of the Skyscrapers, Chicago, Illinois,
January 610, 1986.
21. Irwin, P. A., Ferraro, V., Stone, G. K., Wind Induced
FIRST QUARTER / 1993 15
Motions of Buildings, Proceedings Symposium/Work-
shop on Serviceability of Buildings (Movements, Defor-
mations, Vibrations) Volume I, University of Ottawa, May
1618, 1988.
22. Islam, M. S., Ellingwood, B., Corotis, R. B., Dynamic
Response of Tall Buildings to Stochastic Wind Load,
Journal of Structural Engineering, ASCE, Volume 116,
No. 11, November 1990.
23. Isyumov, N., Poole, M., Wind Induced Torque on Square
and Rectangular Building Shapes, Journal of Wind En-
gineering and Industrial Aerodyamics, 13, 1983.
24. Kareem, A., Lateral Torsional Motion of Tall Buildings
to Wind Loads, Journal of Structural Engineering,
ASCE, Volume 111, No. 11, November 1985.
25. Khan, F., Parmelee, R., Service Criteria for Tall Build-
ings for Wind Loadings, Proceedings 3rd International
Conference on Wind Effects on Buildings and Structures,
Tokyo, Japan, 1971.
26. Minimum Design Loads for Buildings and Other Struc-
tures, ASCE 7-88, American Society of Civil Engineers.
27. Metal Curtain Wall Manual, American Architectural
Manufacturers Association, Des Plaines, Illinois, 1989.
28. Planning and Environmental Criteria for Tall Buildings,
A Monograph on Planning and Design of Tall Buildings,
ASCE, Volume PC, Chapter PC-13, 1981.
29. Standard Building Code, Southern Building Code Con-
gress, 1991 Edition.
30. Structural Design of Tall Steel Buildings, A Monograph
on Planning and Design of Tall Buildings, ASCE, Volume
SB, Chapter SB-5, 1979.
31. Structural Serviceability: A Critical Appraisal of Re-
search Needs, Journal of Structural Engineering, ASCE,
Volume 112, No. 12, December 1986.
32. Supplement to the National Building Code of Canada
1990, National Research Council of Canada, Ottawa,
1990.
33. Tallin, A., Ellingwood, B., Serviceability Limit States:
Wind Induced Vibrations, Journal of Structural Engi-
neering, ASCE, Volume 110, No. 10, October 1984.
34. Brick Veneer Steel Stud Panel Walls, Technical Notes
on Brick Construction No. 28B (revised February 1987),
Brick Institute of America.
35. Uniform Building Code, International Conference of
Building Officials, 1991 Edition.
36. Wind Drift Design of Steel-Framed Buildings: State-of-
the-Art Report, Journal of Structural Engineering,
ASCE, Volume 114, No. 9, September 1988.
37. Vickery, B. J., Isyumov, N., Davenport, A. G., The Role
of Damping, Mass, and Stiffness in the Reduction of Wind
Effects on Structures, Journal of Wind Engineering and
Industrial Aerodynamics, 11, 1983.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 16
INTRODUCTION
In practice, the majority of single angle struts are eccentri-
cally loaded, and an attempt to calculate exactly the strength
of such members is a task of formidable complexity. There-
fore, the designer often relies on approximate methods and
guidelines, in which the key ingredient is the value of the
design strength when the strut is loaded concentrically (AISC
1986, ASCE 1988, and Adluri and Madugula 1992).
In this brief note, a step-by-step solution and load tables
are presented for the determination of the design strength of
concentrically loaded single angle struts according to the
AISC 1986 LRFD Specification. Equation A-E3-7, of Appen-
dix E of the LRFD Specification requires the computation of
the torsional-flexural buckling stress as the smallest of the
three roots of a cubic equation.
GEOMETRY AND COORDINATE SYSTEMS
Consider the unsymmetrical single angle section, shown in
Figure 1, with geomatric axes x and y and principal axes u
and v, each passing through the centroid C. Let be the angle
of inclination between the horizonal axis x and the principal
axis u; and b
1
, b
2
, and t be the dimensions of the vertical leg,
horizontal leg, and thickness of the angle, respectively. The
shear center of the section SC is located at the intersection of
the center lines of the two legs and measures from the centroid
x
o
and y
o
, in the (x, y) system, and u
o
and v
o
, in the (u, v)
system, as shown in Figure 1. They are defined as follows:
x
o
x


t
2
y
o
y


t
2
(1)
u
o
y
o
sin + x
o
cos
v
o
y
o
cos x
o
sin
x

, y

, and Tan tabulated in the AISC Manual (AISC 1986).


FUNDAMENTAL FORMULAE
For an unequal-angle member under concentric loading the
elastic flexural-torsional buckling stress is defined as the
lowest root of the following cubic equation (Timoshenko and
Gere 1961):
(F
e
F
eu
)(F
e
F
ev
)(F
e
F
euz
) F
e

2
(F
e
F
ev
)

u
o
r
o
_

,
2
F
e

2
(F
e
F
eu
)

v
o
r
o
_

,
2
0 (2)
in which
F
eu

2
E

K
u
L
r
u
_

,
2
F
ey

2
E

K
v
L
r
v
_

,
2
F
ez

1
Ar
o
2

2
EC
w
(K
z
L)
2
+ GJ
1
1
]
(3)
In the above equations, F
eu
F
ev
, and F
ez
are the elastic flexural
buckling stresses about the u and v axes and the elastic
torsional buckling stress about the z-axis respectively. K
u
L,
K
v
L are the effective lengths for bending about the principal
axes u and v. K
z
L is the effective length for twisting about the
z-axis, which passes through the centroid of the section and
A. Zureick is Associate Professor, School of Civil Engineering,
Georgia Institute of Technology, Atlanta, GA.
Design Strength of Concentrically Loaded
Single Angle Struts
A. ZUREICK
Figure 1
FIRST QUARTER / 1993 17
coincides with the angles longitudinal axis. The cross-sec-
tional area of the angle is A; C
w
and J are the warping constant
and the St. Venant torsion constant, respectively. E and G are
the modulus of elasticity and the shear modulus of the steel
with values of 29,000 ksi and 11,200 ksi respectively. r
o
, the
polar radius of gyration of the cross section about the shear
center, is defined as
r
o

u
o
2
+ v
o
2
+
I
u
+ I
v
A
(4)
Equation 2 is identical to AISC LRFD Equation A-E3-7
and can be written conveniently in the form
F
e

3
+ a
2
F
e

2
+ a
1
F
e
+ a
0
0 (5)
where
a
2

1
H

1
v
o
2
r
o
2
_

,
F
eu
+

1
u
o
2
r
o
2
_

,
F
ev
+ F
ez

1
1
]

,
a
1

1
H
(F
eu
F
ev
+ F
eu
F
ez
+ F
ev
F
ez
)
,
a
0

1
H
(F
eu
F
ev
F
ez
) (6)
in which H is defined as
H 1
u
o
2
+ v
o
2
r
o
2
(7)
SOLUTION OF THE CUBIC EQUATION
In general, a cubic equation has either three real roots or one
real and two complex conjugate roots. In the case of buckling
of an unsymmetrical section, however, it can be shown that
the three roots of the cubic equation are always real and
positive (See, for example, the proof in texts by Timoshenko
and Gere, 1961 or Galambos, 1968). In such a case the
solution of the cubic equation can most conveniently be
obtained as follows: Let
M
3a
1
a
2
2
9

,
N
9a
2
a
1
2a
2
3
27a
0
54

,
arccos
N
M
3

(8)
then the three real roots are given by
F
e1
2 Mcos

3

a
2
3

,
F
e2
2 Mcos

3
+
2
3
_

,

a
2
3

,
F
e3
2 Mcos

3
+
4
3
_

,

a
2
3
(9)
The above solution was first presented in 1615 by the
French mathematician Franois Vite (Kline 1972) and is
used in most mathematical handbooks nowadays (Tuma
1970, Beyer 1982). In the above equation the quantity M is
always negative.
DESIGN STRENGTH OF CONCENTRICALLY
LOADED STRUTS
Upon solving the cubic equation, the critical flexural-tor-
sional buckling stress F
c
for the angle strut is defined as the
lowest of the three roots obtained previously. Thus,
F
e
min [F
e1
, F
e2
, F
e3
]
from which the slenderness parameter
c
, the nominal critical
stress
c
F
cr
, and the design strength
c
P
n
, can be calculated
according to AISC LRFD Appendix E. A schematic showing
all design steps for calculating the design strength of concen-
trically loaded angle struts is shown in Figure 2. Figure 2
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 18
Example
Calculate design strength for a 9 ft long concentrically loaded
53
1

4
single angle made of A36 steel.
Solution
Given:
K
u
L K
v
L K
z
L 9 ft, F
y
= 36 ksi, E = 29,000 ksi,
G = 11,200 ksi
For L53
1

4
:
A = 1.94 in.
2
, r
x
= 1.62 in., r
y
= 0.861 in., r
v
= 0.663 in.,
Tan = 0.371 or ( = 0.355 rad. = 20.35)
C
w
0.0606 in.
6
, J 0.0438 in.
4
,
Ix 5.11 in.
4
Iy 1.44 in.
4
Properties
I
v
Ar
v
2
(1.94)(0.663)
2
0.852 in.
4
I
u
I
x
+ I
y
I
v
5.11 + 1.44 0.852 5.698 in.
4
r
u

I
u
A

5.698
1.94
1.713 in.
x
o
x


t
2
0.657
0.25
2
0.532 in.
y
o
y


t
2
1.66
0.25
2
1.535 in.
u
o
y
o
sin + x
o
cos 1.535 sin (20.35) +
0.532 cos (20.35) = 1.033 in.
v
o
y
o
cos x
o
sin 1.535 cos (20.35)
0.532 sin (20.35) = 1.254 in.
r
o

u
o
2
+ v
o
2
+
I
u
+ I
v
A

(1.033)
2
+ (1.254)
2
+
5.698 + 0.852
1.94
2.453 in.
H 1
u
o
2
+ v
o
2
r
o
2
1
(1.033)
2
+ (1.254)
2
(2.452)
2
0.561
Check local buckling
b
t

5
1

4
20 >
76
36
12.67 <
155
36
25.83
Thus, use AISC LRFD Eq. A-B5-1
Q
s
1.34 0.00447

b
t
_

,
F
y
1.34 0.00447(20) 36 0.804
Q Q
s
0.804
Check global buckling
The slenderness ratios about the principal axes are
K
u
L
r
u

9(12)
1.713
63.05
,

K
v
L
r
v

9(12)
0.663
162.9
The elastic buckling stresses about the u, v, and z axes are
F
eu

2
E

K
u
L
r
u
_

,
2

2
(29,000)
(63.05)
2
72 ksi,
F
ev

2
E

K
v
L
r
v
_

,
2

2
(29,000)
(162.9)
2
10.78 ksi
F
ez

2
EC
w
(K
z
L)
2
+ GJ
_

,

1
Ar

o
2

2
(29,000)(0.0606)
(108)
2
+(11,200)(0.0438)
_

,

1
(1.94)(2.453)
2

(1.48 + 490)
11.67
42.1 ksi
Now it is required to solve the cubic equation
F
e

3
+ a
2
F
e

2
+ a
1
F
e
+ a
0
0, where
a
2

1
H

1
v
o
2
r
o
2
_

,
F
eu
+

1
u
o
2
r
o
2
_

,
F
ev
+ F
ez

1
1
]

1
0.561

1
(1.254)
2
(2.453)
2
_

,
(72)+

1
(1.033)
2
(2.453)
2
_

,
(10.78)+(42.1)
1
1
]
186
a
1

1
H
(F
eu
F
ev
+ F
eu
F
ez
+ F
ev
F
ez
)
1
0.561
[(72)(10.78)
+ (72)(42.1) + (10.78)(42.1)] 7,596
a
0

1
H
(F
eu
F
ev
F
ez
)
1
0.561
(72)(10.78)(42.1) 58,247
M
3a
1
a
2
2
9

3(7,596) (186)
2
9
1,312
N
9a
1
a
2
27a
0
2a
2
3
54
FIRST QUARTER / 1993 19

9(7,596)(186) 27(58,247) 2(186)
3
54
31,976
arccos
N
M
3

arccos
31,976
(1,312)
3
arccos (0.6729) 0.832 rad.
F
e1
2 M cos

3

a
2
3
2 (1,312) cos
0.832
3

(186)
3
131.6 ksi
F
e2
2 M cos

3
+
2
3
_

,

a
2
3
2 (1,312) cos

0.832
3
+
2
3
_

,

(186)
3
9.98 ksi
F
e3
2 M cos

3
+
4
3
_

,

a
2
3
2 (1,312) cos

0.832
3
+
4
3
_

,

(186)
3
44.4 ksi
Therefore
F
e
= min[131.6, 9.98, 44.4] = 9.98 ksi
(please note how close the flexural-torsional buckling stress
(F
e
= 9.98 ksi) is to that calculated from the flexural buckling
about the minor axis (F
ev
= 10.78 ksi). The difference here is
only eight percent.)

F
y
F
e

36
9.98
1.9

e
Q 1.9 0.804 1.704 > 1.5
F
cr

0.877

e
2
1
1
]
F
y

0.877
(1.9)
2
1
1
]
(36) 8.74 ksi
P
n
F
cr
A
g
0.85(8.74)(1.94) 14.4 kips
Of some note, if the flexural buckling stress with respect to
the minor axis (F
ev
= 10.78 ksi) were used instead of the
flexural-torsional buckling stress (F
e
= 9.98 ksi), the design
strength of the angle member would be 15.6 kips. Such an
increase (8.3 percent) may not be regarded as significant. It
should be emphasized, however, that for a strut with a small
slenderness ratio the difference between the elastic flexural
and the elastic flexural-torsional buckling stresses can be
quite large. A procedure related to the design of axially loaded
compressed angles by the method of minor axis buckling was
presented by Galambos (1991).
DESIGN TABLES
The procedure outlined in this note was implemented to
generate angle load tables for 139 equal and unequal leg angle
sections made of either A36 or Grade 50 steel. Tabulated loads
are for, K
u
L = K
v
L = L, and are terminated when the slender-
ness ratio with respect to the minor axis of the angle exceeds
200. Numerical values are rounded to the nearest whole
number. It should be noted that it is, of course, not necessary
to use the solution of the cubic equation to calculate the design
strength of a concentrically loaded equal leg angle strut.
Equal leg angle struts can be regarded as singly symmetric
sections (symmetry about the u-axis) for which either
flexural-torsional buckling about the axis of symmetry (u-
axis) or flexural buckling about an axis perpendicular to the
axis of symmetry (v-axis) may occur. In such a case the cubic
equation is reduced to a quadratic equation, the solution of
which is given as Equation A-E3-6 in the AISC LRFD Speci-
fication. Solutions from either AISC LRFD Equation A-E3-6
or A-E3-7 yield the same results when applied to equal leg
angles.
REFERENCES
Adluri, S. M. R., and Madugula, M. K. S. (1992). Eccentri-
cally Loaded Steel Single Angle Struts, AISC Engineering
Journal, 2nd Quarter 1992, pp. 5966.
AISC (1986), Manual of Steel ConstructionLoad and Re-
sistance Factor Design, First Edition, Chicago, IL.
ASCE (1988), Guide for Design of Steel Transmission Tow-
ers, Manuals and Reports on Engineering Practice No. 52,
Second Edition, New York.
Beyer, William H. (1982). CRC Mathematical Tables, 26th
Edition, CRC Press, Boca Raton, FL.
Galambos, T. V. (1968). Structural Members and Frames,
Prentice Hall, Englewood Cliffs, NJ.
Galambos, T. V. (1991), Design of Axially Loaded Com-
pressed Angles, Proceedings of the SSRC Annual Technical
Meeting, April 1517, Chicago, pp. 353359.
Kline, Morris, (1972). Mathematical Thought from Ancient
to Modern Times, Oxford University Press, New York.
Timoshenko, S. P., and Gere, J. M. (1961). Theory of Elastic
Stability, Second Edition, McGraw-Hill Book Company.
Tuma, Jan J. (1970). Engineering Mathematics Handbook,
McGraw-Hill Book Company.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 20
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 9 4
Thickness
5

8
9

16
1

2
Wt. / ft 26.3 23.8 21.3
F
y 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 226 291 195 248 164 205
1
2
3
4
5
203
191
182
171
158
254
234
220
204
184
173
161
154
145
135
213
194
183
171
156
142
131
125
119
112
171
155
146
137
127
6
7
8
9
10
143
126
108
92
76
161
137
113
92
76
123
109
95
81
68
138
119
100
82
68
102
92
81
70
59
114
100
85
71
59
11
12
13
14
64
54
46
40
64
54
46
40
57
49
42
36
57
49
42
36
50
43
37
32
50
43
37
32
Size 8 6
Thickness 1
7

8
3

4
5

8
9

16
1

2
7

16
Wt. / ft 44.2 39.1 33.8 28.5 25.7 23.0 20.2
F
y 36 50 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 398 553 352 489 304 422 255 332 222 286 188 239 154 192
1
2
3
4
5
383
376
371
363
350
524
511
502
487
462
335
327
323
317
307
457
442
434
423
405
285
276
272
268
262
385
369
362
354
343
232
222
218
215
211
293
277
271
266
260
198
188
184
182
179
247
231
225
222
217
164
154
151
149
147
201
186
181
178
175
131
121
118
117
115
156
142
138
135
133
6
7
8
9
10
333
314
293
271
248
432
398
362
324
287
293
277
258
239
219
379
350
318
286
253
251
238
223
207
190
324
301
274
247
219
205
196
184
171
158
250
236
218
198
178
175
169
161
150
139
211
201
189
173
157
144
140
135
128
119
171
165
157
146
134
113
111
108
104
98
131
128
123
117
110
11
12
13
14
15
225
203
181
159
139
251
217
185
160
139
199
179
159
141
123
221
191
163
141
123
173
155
139
123
107
192
166
142
123
107
144
130
116
102
90
157
137
119
103
90
128
116
104
93
82
140
124
108
94
82
110
100
91
82
73
121
108
95
83
73
92
85
78
70
63
101
91
82
72
64
16
17
18
19
20
122
109
97
87
79
122
109
97
87
79
108
96
86
77
69
108
96
86
77
69
95
84
75
67
61
95
84
75
67
61
79
70
63
56
51
79
70
63
56
51
72
64
57
52
47
72
64
57
52
47
64
57
51
46
42
64
57
51
46
42
56
50
45
41
37
56
50
45
41
37
21 71 71 63 63 55 55 46 46 42 42 38 38 33 33
y v
x
u
v y
u
x
FIRST QUARTER / 1993 21
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 8 4
Thickness 1
3

4
9

16
1

2
Wt. / ft 37.4 28.7 21.9 19.6
F
y 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 337 468 258 359 189 243 160 204
1
2
3
4
5
323
312
297
277
252
442
421
394
356
313
241
231
221
207
190
327
308
290
264
234
169
159
153
145
134
211
196
186
173
157
140
131
126
120
113
172
158
151
142
130
6
7
8
9
10
224
196
167
140
114
266
220
176
140
114
170
148
127
106
87
200
166
133
107
87
122
108
93
79
66
138
118
98
80
66
103
92
81
69
58
116
101
85
70
58
11
12
13
14
95
80
68
59
95
80
68
59
73
61
53
45
73
61
53
45
55
47
40
35
55
47
40
35
49
42
36
31
49
42
36
31
Size 7 4
Thickness
3

4
5

8
1

2
3

8
Wt. / ft 26.2 22.1 17.9 13.6
F
y 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 235 327 198 272 155 200 102 127
1
2
3
4
5
222
215
207
193
177
302
289
273
249
220
183
176
170
160
147
244
231
220
203
180
138
131
127
121
113
172
162
155
145
132
85
80
77
75
71
101
94
90
86
81
6
7
8
9
10
158
139
119
100
82
189
157
127
101
82
132
116
100
84
69
156
130
106
85
69
102
91
79
67
56
117
100
84
68
56
66
61
54
47
41
74
66
58
49
41
11
12
13
14
68
58
49
42
68
58
49
42
58
49
42
36
58
49
42
36
47
40
43
29
47
40
34
29
35
29
25
22
35
29
25
22
y v
x
u
v y
u
x
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 22
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 6 4
Thickness
7

8
3

4
5

8
9

16
1

2
7

16
3

8
5

16
Wt. / ft 27.2 23.6 20.0 18.1 16.2 14.3 12.3 10.3
F
y 36 50 36 50 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 244 339 212 295 179 249 162 226 145 194 124 161 101 128 76 94
1
2
3
4
5
236
230
220
205
187
324
313
293
266
234
203
198
190
177
162
277
268
253
230
203
168
164
158
149
136
228
220
209
192
170
150
146
141
133
123
202
194
186
172
153
132
127
124
118
109
170
163
156
146
132
110
106
103
99
92
137
130
126
119
109
86
82
80
78
73
105
99
96
92
86
62
59
58
56
54
73
68
66
64
61
6
7
8
9
10
167
146
125
105
86
200
166
134
106
86
145
127
109
91
75
173
144
117
93
75
122
107
92
77
64
146
122
98
78
64
110
97
83
70
58
131
110
89
71
58
98
86
74
63
52
114
96
79
63
52
84
74
64
55
45
96
82
69
56
45
68
61
54
46
39
77
68
57
47
39
51
47
42
37
32
57
51
45
38
32
11
12
13
14
71
60
51
44
71
60
51
44
62
52
45
39
62
52
45
39
53
44
38
33
53
44
38
33
48
40
34
30
48
40
34
30
43
36
31
27
43
36
31
27
38
32
27
24
38
32
27
24
33
28
24
20
33
28
24
20
27
23
20
17
27
23
20
17
Size 6 3
1

2
Thickness
1

2
3

8
5

16
Wt. / ft 15.3 11.7 9.8
F
y 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 138 184 95 121 72 89
1
2
3
4
5
125
120
114
106
95
161
153
143
129
111
82
78
75
71
65
99
93
89
83
74
59
56
54
52
48
70
65
62
59
54
6
7
8
9
10
82
69
57
46
37
92
74
57
46
37
58
50
42
34
28
64
53
43
34
28
44
39
34
28
23
48
42
35
28
23
11
12
31
26
31
26
24
20
24
20
20
17
20
17
y v
x
u
v y
u
x
FIRST QUARTER / 1993 23
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 5 3
1

2
Thickness
3

4
5

8
1

2
7

16
3

8
5

16
1

4
Wt. / ft 19.8 16.8 13.6 12.0 10.4 8.7 7.0
F
y 36 50 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 178 247 151 209 122 170 108 147 92 119 71 91 51 62
1
2
3
4
5
172
166
156
142
126
235
225
206
181
153
143
139
131
120
106
195
188
173
152
129
113
110
105
97
86
153
147
138
123
104
98
95
91
85
76
128
123
117
105
90
80
78
75
71
64
101
96
92
85
74
60
58
56
54
50
73
69
67
63
57
40
38
37
36
34
47
44
43
41
39
6
7
8
9
10
108
91
74
59
48
124
97
75
59
48
92
77
63
50
41
105
83
63
50
41
74
63
51
41
33
85
67
52
41
33
66
55
45
36
30
74
59
46
36
30
56
47
39
32
26
62
50
40
32
26
44
38
32
26
22
49
41
33
26
22
32
28
25
21
17
35
30
26
21
17
11
12
40
33
40
33
34
28
34
28
28
23
28
23
25
21
25
21
21
18
21
18
18
15
18
15
14
12
14
12
Size 5 3
Thickness
5

8
1
2
7
16
3

8
5

16
1
4
Wt. / ft 15.7 12.8 11.3 9.8 8.2 6.6
F
y 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 141 196 115 159 101 138 86 112 67 85 48 58
1
2
3
4
5
134
128
117
103
88
182
171
152
127
101
106
102
94
83
71
143
135
121
102
82
92
88
82
73
62
120
113
103
88
71
75
72
68
61
53
94
89
82
72
59
56
54
51
47
41
68
64
60
54
46
38
36
34
32
29
44
41
39
36
32
6
7
8
9
10
72
57
44
35
28
77
57
44
35
28
58
46
36
28
23
62
46
36
28
23
51
41
32
25
20
55
41
32
25
20
44
35
27
22
18
46
35
27
22
18
35
29
23
18
15
37
29
23
18
15
26
22
18
14
12
28
22
18
14
12
y v
x
u
v y
u
x
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 24
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 4 3
1

2
Thickness
5

8
1

2
7

16
3

8
5

16
1

4
Wt. / ft 14.7 11.9 10.6 9.1 7.7 6.2
F
y 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 132 183 107 149 95 131 82 113 69 89 50 64
1
2
3
4
5
127
124
115
104
91
174
168
152
132
110
101
99
94
85
74
137
134
123
107
89
88
86
82
75
66
118
116
109
95
79
74
72
70
64
57
99
96
92
82
68
59
58
57
54
48
74
72
70
65
56
41
40
39
38
36
49
47
47
45
41
6
7
8
9
10
78
64
51
41
33
88
67
51
41
33
63
52
42
33
27
72
55
42
33
27
56
46
37
30
24
63
49
37
30
24
49
40
32
26
21
55
42
33
26
21
41
34
28
22
18
46
36
28
22
18
31
27
22
18
14
35
28
22
18
14
11
12
27 27 22
19
22
19
20
17
20
17
17
14
17
14
15
12
15
12
12
10
12
10
Size 4 3
Thickness
5

8
1

2
7

16
3

8
5

16
1

4
Wt. / ft 13.6 11.1 9.8 8.5 7.2 5.8
F
y 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L

0 122 169 99 138 88 122 76 105 64 83 47 60
1
2
3
4
5
117
112
102
90
76
161
151
133
111
88
94
91
83
73
62
128
122
108
90
72
82
79
73
65
55
111
106
95
80
63
69
67
63
56
47
92
89
81
68
55
56
54
52
46
40
70
67
63
55
45
39
38
37
34
30
47
45
43
40
34
6
7
8
9
10
62
49
37
29
24
66
49
37
29
24
51
40
31
24
20
54
40
31
24
20
45
35
27
21
17
48
35
27
21
17
39
31
24
19
15
41
31
24
19
15
33
26
20
16
13
35
26
20
16
13
25
21
16
13
10
27
21
16
13
10
Size 3
1

2
3
Thickness
1

2
7

16
3

8
5

16
1

4
Wt. / ft 10.2 9.1 7.9 6.6 5.4
F
y 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L 0 92 128 81 113 70 98 59 81 46 60
1
2
3
4
5
88
84
77
67
56
120
114
99
82
64
76
74
68
59
50
104
100
88
73
57
65
64
59
51
43
88
85
76
63
50
53
52
49
43
36
69
67
62
52
41
39
38
37
34
29
48
46
45
40
32
6
7
8
9
10
45
35
27
21
17
47
35
27
21
17
40
31
24
19
15
42
31
24
19
15
35
27
21
16
13
37
27
21
16
13
29
23
17
14
11
31
23
17
14
11
24
19
14
11
9
25
19
14
11
9
y v
x
u
v y
u
x
FIRST QUARTER / 1993 25
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 3
1
2 2
1
2
Thickness
1
2
7
16
3
8
5
16
1
4
Wt. / ft 9.4 8.3 7.2 6.1 4.9
F
y 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 84 117 74 103 65 90 54 75 43 55
1
2
3
4
5
6
7
8
9
80
75
66
55
43
32
24
18
109
99
83
64
46
32
24
18
70
66
58
48
38
28
21
16
95
87
73
57
41
28
21
16
60
57
50
42
33
25
18
14
80
75
63
49
35
25
18
14
49
47
42
35
28
21
15
12
9
64
60
52
41
30
21
15
12
9
36
35
32
28
22
17
13
10
8
44
42
38
31
24
17
13
10
8
Size 3 2
1
2
Thickness
1
2
7
16
3
8
5
16
1
4
3
16
Wt. / ft 8.5 7.6 6.6 5.6 4.5 3.39
F
y 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,
k
L

0 77 106 68 94 59 82 50 69 40 53 28 35
1
2
3
4
5
6
7
8
74
68
59
49
38
28
20
16
101
91
75
57
40
28
20
16
65
60
52
43
34
25
18
14
88
80
66
50
35
25
18
14
55
52
46
38
29
21
16
12
75
69
57
44
31
21
16
12
45
44
38
32
25
18
13
10
61
58
48
37
26
18
13
10
35
34
31
26
20
15
11
8
45
43
38
30
21
15
11
8
23
22
21
18
15
11
8
6
27
26
25
21
16
11
8
6
Size 3 2
Thickness
1
2
7
16
3
8
5
16
1
4
3
16
Wt. / ft 7.7 6.8 5.9 5.0 4.1 3.07
F
y 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,
k
L

0 69 96 61 85 53 74 45 62 36 49 25 32
1
2
3
4
5
6
7
65
58
47
35
24
17
12
89
75
56
38
24
17
12
58
51
42
31
22
15
11
78
67
50
34
22
15
11
49
44
36
27
19
13
10
67
57
43
29
19
13
10
41
37
30
23
16
11
8
55
48
36
25
16
11
8
32
30
25
19
13
9
7
41
37
29
20
13
9
7
21
20
17
13
10
7
5
25
23
19
14
10
7
5
Size 2
1
2 2
Thickness
3
8
5
16
1
4
3
16
Wt. / ft 5.3 4.5 3.62 2.75
F
y 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,
k
L

0 47 66 40 56 32 45 24 32
1
2
3
4
5
6
7
45
40
32
24
16
11
8
61
52
38
25
16
11
8
37
34
27
20
14
10
7
51
44
32
21
14
10
7
29
27
22
16
11
8
6
39
35
26
17
11
8
6
21
20
16
12
9
6
4
25
24
19
13
9
6
4
y v
x
u
v y
u
x
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 26
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 8 8
Thickness 1
1

8 1
7

8
3

4
5

8
9

16
1

2
Wt. / ft 56.9 51.0 45.0 38.9 32.7 29.6 26.4
F
y 36 50 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 511 710 459 638 404 561 349 485 293 382 255 328 216 275
5 473 637 423 569 364 486 304 400 241 296 202 244 164 193
6
7
8
9
10
457
439
419
397
374
607
574
538
500
460
410
394
376
357
336
546
516
483
449
414
362
347
332
315
297
481
455
427
397
366
302
300
287
273
257
397
393
370
344
318
239
238
237
229
217
294
291
289
278
258
201
200
198
197
191
242
240
238
236
227
163
162
161
160
159
192
190
189
188
186
11
12
13
14
15
351
326
302
278
254
420
381
342
304
268
315
293
271
249
228
378
342
307
273
240
278
259
240
221
202
335
303
273
243
214
242
225
209
192
176
291
264
238
212
187
203
190
176
162
148
237
216
196
176
157
180
168
157
145
133
210
193
176
159
143
155
146
136
126
117
180
167
153
139
126
16
17
18
19
20
230
208
186
167
151
235
208
186
167
151
207
187
167
150
135
211
187
167
150
135
184
166
149
134
121
188
167
149
134
121
160
145
130
117
105
165
146
130
117
105
135
122
110
98
89
139
123
110
98
89
122
111
100
90
81
127
113
100
90
81
107
98
89
80
73
113
100
90
80
73
21
22
23
24
25
137
124
114
105
96
137
124
114
105
96
123
112
102
94
87
123
112
102
94
87
109
100
91
84
77
109
100
91
84
77
96
87
80
73
67
96
87
80
73
67
81
73
67
62
57
81
73
67
62
57
74
67
61
56
52
74
67
61
56
52
66
60
55
50
46
66
60
55
50
46
26 89 89 80 80 71 71 62 62 53 53 48 48 43 43
x
u
y
v
x
u
y
v
FIRST QUARTER / 1993 27
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 6 6
Thickness 1
7

8
3

4
5

8
9

16
1

2
7

16
3

8
5

16
Wt. / ft 37.4 33.1 28.7 24.2 21.9 19.6 17.2 14.9 12.4
F
y 36 50 36 50 36 50 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 337 468 298 414 258 359 218 302 197 273 176 235 151 195 122 155 92 114
1
2
3
4
5
327
324
320
308
293
449
443
436
413
368
287
283
281
272
259
392
385
382
366
341
245
241
239
236
225
334
326
322
317
296
202
197
195
194
190
272
263
260
258
250
179
174
172
171
170
240
230
227
225
223
157
151
149
148
147
201
191
188
186
184
130
124
122
121
120
161
151
148
147
145
101
95
93
92
92
122
113
110
109
108
72
67
65
65
64
84
76
74
73
73
6
7
8
9
10
276
257
236
215
193
354
321
286
251
217
244
227
209
190
171
314
284
253
222
192
212
197
181
165
148
272
246
219
192
166
179
167
154
140
126
230
209
186
164
142
162
151
139
127
114
208
189
168
148
128
145
135
124
113
102
181
164
147
130
114
119
117
108
99
89
144
140
127
113
99
91
90
89
82
75
107
106
104
94
83
64
63
63
62
60
72
71
71
70
67
11
12
13
14
15
172
152
132
114
99
184
155
132
114
99
152
134
117
101
88
163
137
117
101
88
132
116
101
87
76
141
119
101
87
76
113
99
87
75
65
121
102
87
75
65
102
90
78
68
59
109
92
78
68
59
91
80
70
61
53
97
82
70
61
53
80
71
62
54
47
86
74
63
54
47
67
60
53
47
41
73
63
54
47
41
54
49
44
39
35
59
53
46
40
35
16
17
18
19
87
77
69
62
87
77
69
62
77
68
61
55
77
68
61
55
67
59
53
47
67
59
53
47
57
51
45
41
57
51
45
41
52
46
41
37
52
46
41
37
46
41
37
33
46
41
37
33
41
37
33
29
41
37
33
29
36
32
28
25
36
32
28
25
30
27
24
22
30
27
24
22
Size 5 5
Thickness
7

8
3

4
5

8
1

2
7

16
3

8
5

16
Wt. / ft 27.2 23.6 20.0 16.2 14.3 12.3 10.3
F
y 36 50 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 244 339 212 295 179 249 145 202 128 174 109 141 84 107
1
2
3
4
5
237
235
227
215
200
326
322
307
284
257
204
202
198
187
174
279
275
267
247
224
169
167
166
158
147
230
226
223
209
189
133
130
129
128
119
178
173
171
169
154
114
111
110
109
105
149
143
141
140
133
93
90
89
88
87
115
110
108
107
106
69
66
65
64
64
83
78
77
76
75
6
7
8
9
10
183
165
146
128
110
227
197
166
138
112
159
144
127
111
96
198
171
145
120
98
135
122
108
94
81
168
145
123
102
83
110
99
88
77
66
136
118
101
84
68
97
87
78
68
59
119
103
88
74
60
83
75
67
59
51
99
87
75
63
52
63
60
54
48
42
74
69
61
52
44
11
12
13
14
15
93
78
66
57
50
93
78
66
57
50
81
68
58
50
43
81
68
58
50
43
69
58
49
42
37
59
58
49
42
37
56
47
40
35
30
56
47
40
35
30
50
42
36
31
27
50
42
36
31
27
43
36
31
27
23
43
36
31
27
23
36
31
26
23
20
37
31
26
23
20
16 44 44 38 38 32 32 27 27 24 24 20 20 17 17
x
u
y
v
x
u
y
v
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 28
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 4 4
Thickness
3

4
5

8
1

2
7

16
3

8
5

16
1

4
Wt. / ft 18.5 15.7 12.8 11.3 9.8 8.2 6.6
F
y 36 50 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 166 231 141 196 115 159 101 141 88 122 73 95 54 69
1
2
3
4
5
162
158
149
136
122
222
216
198
175
150
136
134
126
116
103
185
183
168
148
127
108
106
103
94
84
146
144
136
121
104
93
92
91
83
74
126
123
121
107
92
78
77
76
72
64
104
101
100
93
80
62
61
60
59
54
78
75
74
72
64
43
42
41
41
40
52
49
49
48
47
6
7
8
9
10
106
90
75
60
49
124
99
76
60
49
90
76
63
51
41
105
84
65
51
41
73
63
52
42
34
86
69
53
42
34
65
55
46
37
30
76
61
47
37
30
56
48
40
32
26
66
53
41
32
26
47
41
34
27
22
54
44
35
27
22
36
32
27
22
18
42
35
28
22
18
11
12
13
40
34
40
34
34
29
34
29
28
24
20
28
24
20
25
21
18
25
21
18
22
18
16
22
18
16
18
15
13
18
15
13
15
13
11
15
13
11
Size 3
1

2
3
1

2
Thickness
1

2
7

16
3

8
5

16
1

4
Wt. / ft 11.1 9.8 8.5 7.2 5.8
F
y 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L
0 99 138 88 122 76 105 64 88 50 64
1
2
3
4
5
95
93
86
77
66
129
126
113
96
79
82
81
76
68
59
111
110
100
85
69
70
68
66
59
51
93
91
86
74
60
56
55
55
50
43
74
72
71
62
51
41
40
40
39
34
50
49
48
47
39
6
7
8
9
10
55
45
35
28
22
61
46
35
28
22
49
40
31
25
20
54
41
31
25
20
43
35
27
21
17
47
35
27
21
17
36
29
23
18
15
40
30
23
18
15
29
24
19
15
12
32
25
19
15
12
11 19 19 16 16 14 14 12 12 10 10
Size 3 3
Thickness
1

2
7

16
3

8
5

16
1

4
3

16
Wt. / ft 9.4 8.3 7.2 6.1 4.9 3.71
F
y 36 50 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L 0 84 117 74 103 65 90 54 76 44 59 30 39
1
2
3
4
5
81
77
69
59
48
111
103
89
71
54
71
68
61
52
43
96
91
78
63
48
60
59
53
45
37
82
79
68
55
42
49
49
45
38
32
66
65
58
47
35
38
37
36
31
26
48
47
45
37
29
24
23
23
22
19
28
27
27
26
21
6
7
8
9
38
28
22
17
39
28
22
17
33
25
19
15
34
25
19
15
29
22
17
13
30
22
17
13
25
19
14
11
25
19
14
11
20
15
12
9
21
15
12
9
15
12
9
7
16
12
9
7
x
u
y
v
x
u
y
v
FIRST QUARTER / 1993 29
F
y
= 36 ksi
F
y
= 50 ksi
CONCENTRICALLY LOADED
COLUMNS
Single Angles
Design axial strength in kips ( = 0.85)
Size 2
1

2
2
1

2
Thickness
1

2
3

8
5

16
1

4
3

16
Wt. / ft 7.7 5.9 5.0 4.1 3.07
F
y 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,

k
L 0 69 96 53 74 45 62 36 51 27 35
1
2
3
4
5
67
51
52
41
31
91
80
64
47
32
50
47
40
32
24
69
62
49
36
24
42
39
34
27
20
56
52
42
31
21
32
32
27
22
17
43
42
34
25
17
23
22
21
17
13
28
27
25
19
13
6
7
8
22
16
12
22
16
12
17
12
9
17
12
9
14
11
8
14
11
8
12
9
7
12
9
7
9
7
5
9
7
5
Size 2 2
Thickness
3

8
5

16
1

4
3

16
1

8
Wt. / ft 4.7 3.92 3.19 2.44 1.65
F
y 36 50 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n

f
e
e
t
,
k
L
0 42 58 35 49 29 40 22 30 13 17
1
2
3
4
5
40
34
27
19
12
54
44
31
19
12
33
29
22
16
10
46
37
26
16
10
27
24
18
13
8
36
30
21
13
8
19
18
14
10
7
25
23
17
10
7
10
10
9
7
5
12
12
10
7
5
6 8 8 7 7 6 6 5 5 3 3
Size 1
3

4
1
3

4
1
1

2
1
1

2
Thickness
1

4
3

16
1

4
3

16
Wt. / ft 2.77 2.12 2.34 1.8
F
y 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h

i
n
f
e
e
t
,

k
L
0 25 35 19 26 21 29 16 22
1
2
3
4
5
23
19
14
9
6
32
24
15
9
6
17
15
11
7
4
23
19
12
7
4
19
15
10
5
26
18
10
5
15
11
7
4
20
14
7
4
Size 1
1

4
1
1

4
1
1

8
1
1

8
1 1
Thickness
1

4
3

16
1

8
1

8
Wt. / ft 1.92 1.48 0.9 0.8
F
y 36 50 36 50 36 50 36 50
E
f
f
e
c
t
i
v
e

l
e
n
g
t
h
i
n

f
e
e
t
,

k
L
0 17 24 13 19 8 11 7 10
1
2
3
4
15
10
6
3
20
12
6
3
12
8
4
2
16
9
4
2
7
4
2
9
5
2
6
3
1
8
3
1
x
u
y
v
x
u
y
v
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 30
It is common practice in crane runway beams to place a
channel, open-side down, over the top flange of a W-section,
as shown in Figure 1, to increase its lateral stability. This is
done because it is not always convenient to brace the com-
pression flange between columns.
The lateral-torsional buckling capacity of a singly symmet-
ric section may be determined by the formulas (Footnote c on
page 6-96 of the current LRFD manual
1
) without knowing the
warping section constant. However, these formulas were
developed from research on three-plate monosymmetric sec-
tions.
2,3,4,5,6,7
If one is to develop a similar equation for the case
of the W-section with a channel cap, it is necessary to return
to the basic theory of lateral-torsional buckling and that
means that one must calculate the warping constant (C
wc
) for
the cross-section.
It is not a difficult matter to do this with a computer, but
since the program may not be available to everyone, it would
be useful to develop a simple empirical equation for prelimi-
nary design, making use of the section properties already
given in the AISC steel manual.
According to Kitipornchai and Trahair,
3
Equation 1, which
provides for calculating the warping section constant, is
exact for an unlipped section and approximate for a lipped
section. Their lipped section of Figure 2 closely approxi-
mates the subject of this paper, the W-section with a channel
cap.
I
w
a
2
I
yc
+ b
2
I
yt
(1)
where
a (1 )h b h
h h
c
+ e I
yc
/ I
y
e (D
L
2
B
c
2
T
L
) / (4I
y
)
I
w
= warping section constant
I
y
= moment of inertia of the combined section about the
axis parallel to web
I
yc
= moment of inertia of the compression flange about
the axis parallel to web
I
yt
= moment of inertia of the tension flange about the axis
parallel to web
h
c
= distance from the centroid of the compression flange
to the centroid of the tension flange
h = distance from the shear center of the compression
flange to the centroid of the tension flange
e = distance from the shear center of the compression
flange to the centroid of the compression flange
D
L
= the depth of the lip
B
c
= the width of the lipped flange
T
L
= the thickness of the lip
Tony Lue is Graduate Assistant, University of Florida, Gaines-
ville, FL.
Duane S. Ellifritt is Crom Professor, University of Forida,
Gainesville, FL.
The Warping Constant for the W-Section
with a Channel Cap
TONY LUE AND DUANE S. ELLIFRITT
Fig. 2. W-section with lipped flange.
Fig. 1. Crane runway beam.
FIRST QUARTER / 1993 31
The authors have enhanced a computer program, which
was originally written in BASIC language by Professor Theo-
dore V. Galambos of the University of Minnesota in Minnea-
polis and converted to FORTRAN language by Dr. Thomas
Sputo of the University of Florida, to calculate the exact
values of warping section constants (C
wc
) of the W-section
with a channel cap. The user need only input the W-section
and channel dimensions. The 28 combined sections shown on
pages 1-98 and 1-99 of the AISC LRFD steel manual are
shown in Table 1 with their warping section constants (C
wc
)
as calculated by the program described.
The ratios of C
wc
/ C
w
are plotted against the ratios of
A
c
/ A
w
, as shown in Figure 3 for the 28 combined sections,
where C
w
, A
c
, and A
w
are explained in Table 1.
When Equation 1 was applied to the case of the W-section
with a channel cap and the results I
w
/ C
w
were compared with
the C
wc
/ C
w
values obtained from the above-mentioned pro-
gram, as shown in Figure 4, it can be observed that Equation 1
gives a very conservative estimate (7 percent to 17 percent)
for this case (the W-section with a channel cap). The equation
also has the added disadvantage of requiring the user to
calculate certain section properties and parameters (I
y
, I
yt
, I
yc
,
, h
c
, h, e, a, and b) first.
The purpose of this paper is, therefore, to present a reason-
ably accurate method of calculating the warping section
constant (C
wc
), using a simple model which can be expressed
by Equations 2, 3, or 4 and known section properties that can
be found in the AISC steel manual.
The proposed method which seemed to offer the most
promise was to use the ratio of the channel area (A
c
) to the
W-section area (A
w
) as the independent variable and plot it
against the ratio of the warping section constant (C
wc
) for the
combined section to the warping section constant (C
w
) for the
W-section alone.
Table 1.
W C or MC
A
c
A
w
C
wc
C
w
A
c
/ A
w
C
wc
/ C
w
36150
33141
2484
36150
33118
30116
33141
2468
2168
2162
3099
2794
33118
30116
2784
2484
1850
3099
2468
2168
1430
2162
1636
1226
1850
1430
1226
1636
1533.9
1533.9
1220.7
*1842.7
1533.9
1533.9
*1842.7
1220.7
1220.7
1220.7
1533.9
1533.9
*1842.7
*1842.7
1533.9
1533.9
1220.7
*1842.7
1533.9
1533.9
1015.3
1533.9
1220.7
1015.3
1533.9
1220.7
1220.7
1533.9
9.96
9.96
6.09
12.60
9.96
9.96
12.60
6.09
6.09
6.09
9.96
9.96
12.60
12.60
9.96
9.96
6.09
12.60
9.96
9.96
4.49
9.96
6.09
4.49
9.96
6.09
6.09
9.96
44.20
41.60
24.70
44.20
34.70
34.20
41.60
20.10
20.00
18.30
29.10
27.70
34.70
34.20
24.80
24.70
14.70
29.10
20.10
20.00
8.85
18.30
10.60
7.65
14.70
8.85
7.65
10.60
146,275
117,627
24,215
160,492
91,050
68,520
128,892
18,386
13,753
12,272
53,937
43,840
98,928
74,319
37,518
28,050
6,746
58,101
20,998
15,750
2,024
13,990
3,477
1,459
7,711
2,255
1,645
4,058
82,200
64,400
12,800
82,200
48,300
34,900
64,400
9,430
6,760
5,960
26,800
21,300
48,300
34,900
17,900
12,800
3,040
26,800
9,430
6,760
887
5,960
1,460
607
3,040
887
607
1,460
0.2250
0.2390
0.2470
0.2850
0.2870
0.2910
0.3029
0.3030
0.3050
0.3330
0.3420
0.3600
0.3630
0.3680
0.4020
0.4030
0.4140
0.4330
0.4960
0.4980
0.5070
0.5440
0.5740
0.5870
0.6780
0.6880
0.7960
0.9400
1.780
1.827
1.892
1.952
1.885
1.963
2.001
1.950
2.034
2.059
2.013
2.058
2.048
2.129
2.096
2.191
2.219
2.168
2.227
2.330
2.282
2.347
2.382
2.404
2.537
2.542
2.710
2.779
WW-section; CChannel; MCMiscellaneous Channel
Ac Area of channel, in.
2
; Aw Area of W-section, in.
2
Cwc Warping section constant (in.
6
) for the W-section with a channel cap, which is obtained by using the program as mentioned in this paper
Cw Warping section constant (in.
6
) for the W-section alone, which can be found in the AISC steel manual
*MC
Metric Conversion:
To convert to Multiply by
inches(in.) millimeters(mm) 25.4
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 32
Several curves (or models) were then fit to these data which
are shown in Figure 3. These curves are represented by the
following equations.
C
wc
C
w

1.31 + 2.55

A
c
A
w
_

,
1.31

A
c
A
w
_

,
2
+ 0.29

A
c
A
w
_

,
3
1
1
]
(2)
C
wc
C
w

1.25 + 2.55

A
c
A
w
_

,
1.31

A
c
A
w
_

,
2
+ 0.29

A
c
A
w
_

,
3
1
1
]
(3)
C
wc
C
w

1.7
A
c
A
w
+ 1
1
1
]

A
c
A
w
(4)
Equations 2, 3, and 4 were superimposed on the data of
Figure 3, and plotted in Figure 5. Equation 2 results in an
estimated error of 2.9 percent to +2.9 percent, and Equa-
tion 3 with an estimated error of 5.9 percent to 0. Equation
4 is a simplified model with fewer terms involved, with an
estimated error of 7.1 percent to +1.5 percent. Equations 2
and 3 were derived based on the multiple linear regression
technique of statistics using the data of Figure 3. Equation 4
was also based on the same statistical technique with some
further modifications.
Equation 2 is the best fit curve for the data of Figure 3, but
does overestimate the warping section constant (C
wc
) by as
much as 2.9 percent. Equation 3 is simply Equation 2 shifted
down until all the data points fall above the curve and may be
up to 5.9 percent conservative. However, for those who want
a formula that they can easily memorize and still get a
conservative result within 7.1 percent, Equation 4 is offered.
The Equations 2, 3, and 4 proposed by the authors require
no calculation of certain section properties and parameters,
and are close to an exact solution. The required parameters
for these equations are the channel area (A
c
), the area of the
W-section (A
w
), and its warping section constant (C
w
), all of
which can be found in the AISC steel manual.
REFERENCES
1. American Institute of Steel Construction, Load and Resis-
tance Factor Design, First Edition, Chicago, Illinois, 1986.
2. Anderson, J. M., Trahair, N. S., Stability of Monosymmet-
ric Beams and Cantilevers, Journal of the Structural
Division, ASCE, Vol. 98, No. ST1, January 1972.
3. Kitipornchai, S., Trahair, N. S., Buckling Properties of
Monosymmetric I-Beams, Journal of the Structural Divi-
sion, ASCE, Vol. 106, No. ST5, May 1980.
4. Kitipornchai, S., Wang, C. M., Trahair, N. S., Buckling of
Monosymmetric I-Beams under Moment Gradient, Jour-
nal of the Structural Division, ASCE, Vol. 112, No. 4, April
1986.
5. Bradford, M. A., Cuk, P. E., Elastic Buckling of Tapered
Monosymmetric I-Beams, Journal of the Structural Divi-
sion, ASCE, Vol. 114, No. 5, May 1988.
6. Galambos, T. V., Guide to Stability Design Criteria for
Metal Structures (SSRC), Wiley, New York, 1988.
7. Trahair, N. S., Bradford, M. A., The Behaviour and Design
of Steel Structures, Chapman and Hall, London, 1988.
Fig. 3. Plot of C
wc
/ C
w
vs. A
c
/ A
w
. Fig. 5. Equations 2, 3, and 4 fit to the data of Figure 3.
Fig. 4. I
w
/ C
w
compared with the data of Figure 3
C
wc
is exact for the W-section with a channel cap.
I
w
is an approximation from Reference 3 for the
W-section with a channel cap (see Equation 1).
FIRST QUARTER / 1993 33
INTRODUCTION
An accurate and complete design will result in an economi-
cal and safe connection. Yet it is entirely common for the
engineer of record to withhold, either intentionally or unin-
tentionally, the information necessary to the fabricator or
detailer to perform a design which is both accurate and
complete. Specifically, actual reactions are seldom shown on
the contract drawings from which the connections must be
designed.
1
AISC states, For economical connections, beam reactions
should be shown on the contract drawings. If these reactions
are not shown, connections must be selected to support one-
half the total uniform load capacityfor the given beam,
span, and grade of steel specified.
2,3
No quantification is
given, however, of the actual difference in economy between
the two cases. In fact, this difference is somewhat elusive as
it may vary greatly among specific examples. For the general
case, however, it is possible to determine a reasonable esti-
mate of the economic sacrifice incurred when a larger con-
nection than required is used. The focus of this paper, then, is
this economic sacrifice. For simplicity, a standard configura-
tion
4
of the double angle connection will be considered in
which only n, the number of bolt rows (and consequently, the
length of the angles), varies. Based on values of n from 2 to
10, the cost of these standard connections will be estimated.
Ranges of n compatible with each beam size group will be
identified and the percent increase in cost which results when
a larger connection than required is used will then be deter-
mined over these ranges.
Note that the practices which routinely result in uneconom-
ical connection designs are not specifically addressed in this
paper. For a discussion of these practices, the reader is re-
ferred to Eliminating the Guesswork in Connection De-
signCommunication of Design Requirements Between
Fabricator and Engineer is Crucial for a Safe and Economic
Structure by W.A. Thornton, in the June 1992 issue of
Modern Steel Construction. Also not addressed is the effect
of standardization on the detailing costs, ease of fabrication,
and overall quality of the constructed product. While in
general, standardization will reduce detailing costs, increase
the ease of fabrication, and lead to improvements in quality
because of decreased variability, these considerations are
more project related than connection specific. Thus, it would
not be feasible to consider their effect in this paper.
THE STANDARD CONFIGURATION
The standard parameters of the double angle connection to be
considered are as follows. The shop and field bolts will be
3

4
-in. diameter A325-N at 3-in. spacing with 1
1

4
-in. edge
distance. The holes will be short-slotted in the outstanding
angle legs (those connected to the supporting member) and
standard otherwise. The angles will be 2L 43
1

16
(SLBB).
This standard configuration produces nine connections with
the number of bolts rows n ranging from 2 to 10. While these
connections will not satisfy every case, they will be adequate
for the typical case and, therefore, will be used in this cost
comparison.
CONNECTION COSTS
The costs which will be considered in this paper can be
divided into three categories: material, shop labor, and field
labor. The material costs include the cost of the bolts, washers
and nuts, and the framing angles. The shop labor cost includes
shearing and punching the angles, punching the supported
and supporting members, and installing the shop bolts. The
field labor cost is comprised of installing the field bolts. While
material costs are readily available, labor costs are seldom a
matter of public knowledge. Furthermore, labor costs will
vary from fabricator to fabricator and from region to region.
Consequently, those presented in this paper should be re-
garded as an average estimate and should by no means be
construed to be universal. The fabricators costs which will
be used in this paper are as follows:
A325 tension control (TC) bolts $90.00 per 100 lbs.
L43
1

16
$14.75 per 100 lbs.
Shop Labor $20.00 per hour
Field Labor $30.00 per hour
These are base costs; selling costs, which would include
overhead and profit, would be higher.
The material and labor costs for double angle connections
Charles J. Carter is staff engineer-structures, The American
Institute of Steel Construction, Inc., Chicago, IL.
Louis F. Geschwindner is professor of architectural engineer-
ing, The Pennsylvania State University, University Park, PA.
The Economic Impact of Overspecifying
Simple Connections
CHARLES J. CARTER AND LOUIS F. GESCHWINDNER
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 34
in the standard configuration are summarized in Table 1. The
bolt material cost is based on a bolt length of 3-in. with one
washer and nut each; about 83 pounds per 100 count. The
angle material cost is based on an angle size of 43
1

16
which weighs 7.7 pounds per foot. The labor costs are based
on the labor time estimates
4
summarized in the same table.
Note that, in each row of bolts, there are three bolts: one shop
bolt and two field bolts. Total costs have been rounded to the
nearest whole dollar.
COMPATIBILITY WITH BEAM SHAPES
The deepest compatible standard connection must fit within
the T-dimension of the beam as listed in Part 1 of the ASD
and LRFD Manuals. As recommended in Part 4 of the ASD
Manual and Part 5 of the LRFD Manual, the depth of the
minimum standard connection should be greater than T/2.
Given these limits, the compatibility of the nine standard
connections with W-shapes is summarized in Table 2. Note
that limitations such as coping, which may further restrict the
maximum value of n are not considered.
PERCENT INCREASE IN CONNECTION COST
Given the allowable variations in n of Table 2, percent in-
creases in connection cost per unnecessary row of bolts
provided are listed in Figure 1. Cells below the heavy line fall
outside the spacially permissible variations in n given in
Table 2. As an example of the use of Figure 1, consider a
W1850 and assume an end reaction which would require
four rows of bolts. Using five rows of bolts instead, the largest
n possible given the T-dimension of a W18, would increase
the connection cost by 26 percent. Similarly, using six rows
of bolts in a W2144 requiring only five rows would result
in a cost increase of 17 percent.
Some general observations may be made from Figure 1.
The predicted range of economic sacrifice when all beams
Table 1.
Material, Shop Labor, Field Labor, and Total Cost Estimates
of Double Angle Connections in Standard Configuration
n L
Bolt
Material
Cost
Angle
Material
Cost
Shop
Labor
Time
Shop
Labor
Cost
Field
Labor
Time
Field
Labor
Cost
Rounded
Total
Cost
10 29
1

2
-in. $22.50 $5.61 0.65 $13.00 2.65 $79.50 $121
9 26
1

2
-in. $20.25 $5.04 0.60 $12.00 2.40 $72.00 $109
8 23
1

2
-in. $18.00 $4.47 0.55 $11.00 2.15 $64.50 $ 98
7 20
1

2
-in. $15.75 $3.90 0.50 $10.00 1.90 $57.00 $ 87
6 17
1

2
-in. $13.50 $3.33 0.45 $9.00 1.60 $48.00 $ 74
5 14
1

2
-in. $11.25 $2.76 0.40 $8.00 1.35 $40.50 $ 63
4 11
1

2
-in. $9.00 $2.19 0.35 $7.00 1.05 $31.50 $ 50
3 8
1

2
-in. $6.75 $1.62 0.30 $6.00 0.80 $24.00 $ 38
2 5
1

2
-in. $4.50 $1.05 0.25 $5.00 0.55 $16.50 $ 27
Table 2.
Compatibility of W-shape Beams
and Standard Connection Depths
Shape Weight
n
min
n
max
W36 359-230
256-135
6
6
10
10
W33 354-118 6 10
W30 235-99 5 9
W27 217-84 5 8
W24 176-55 4 7
W21 166-44 4 6
W18 143-35 3 5
W16 100-26 3 4
W14 132-90
82-43
38-22
3
2
2
3
3
4
W12 87-40
35-14
2
2
3
3
W10 112-33
30-12
2
2
2
3
W8 67-24
21-10
2
2
2
2
FIRST QUARTER / 1993 35
and ranges of n are considered is from 11 percent to 85
percent. As the size of the beam being connected decreases,
the percent change in cost increases. Additionally, and obvi-
ously, as the number of unneccesary bolt rows increases, so
does the percent change in cost increase.
Focusing on the range of typical simple beam sizes, the
variation in percent change can be narrowed. First, do not
consider beams larger than a W24; from Table 2, this would
eliminate the 8, 9, and 10 row connections. Additionally,
consider only uniformly loaded cases, the designs of which
are usually controlled by moment. The actual end reactions
will likely be close to, but still less than, the end reactions
calculated from the Design Uniform Load Tables.
4
Thus, in
most cases, the number of unnecessary rows of bolts will be
one. Accordingly, the cost increase given these limitations is
between 13 percent and 41 percent.
It should not be forgotten, however, that there are many
cases which cannot be classified as typical. Shorter span
beams, often sized for convenience or for similarity to other
beams, and in-fill beams, which may serve no other purpose
than to reduce the unbraced length of another member, may
sustain actual reactions which are significantly lower than
one-half the the total uniform load capacity of the beam. A
similar situation is found in beams controlled by deflection
considerations. In these cases, the percent increase in econ-
omy can be much higher than the range identified as typical.
CONCLUSIONS
A generalized and simplified approach has been taken to
estimate the added cost of providing more rows of bolts in a
simple connection than were necessary. This approach was
intended to estimate the possible economic implications
when the engineer of record does not indicate the actual
reactions for which the connections must be designed on the
contract drawings. While this approach centered on the shop
and field bolted double angle connection, this information is
likely similar to that which might be obtained when other
types of simple shear connections are considered. Given the
potential for unnecessary increase in connection cost, the
engineer of record should always indicate the actual reactions
on the contract drawings. In this manner, the best opportunity
for safe and economical connections will be realized.
REFERENCES
1. Thornton, W. A., Eliminating the Guesswork in Connec-
tion DesignCommunication of Design Requirements
Between Fabricator and Engineer is Crucial for a Safe and
Economic Structure, Modern Steel Construction, June
1992, p. 27.
2. American Institute of Steel Construction, Manual of Steel
Construction, Allowable Stress Design, 9th Ed., Chicago,
IL, 1989, pp. 49.
3. American Institute of Steel Construction, Manual of Steel
Construction, Load and Resistance Factor Design, 1st Ed.,
Chicago, IL, 1986, pp. 515.
4. Carter, Charles J., Standardizing Simple Shear Connec-
tions in Steel, Master of Science Thesis, Pennsylvania
State University, University Park, PA, 1991.
Fig.1. Percent increase in connection cost.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 36
The approximate solutions for K-factor in Pierre Dumonteils
paper will assist engineers in programming both spreadsheet
and general language column solutions.
While the general equation for K-factor is rapidly solved
by Newtons Method of Successive Approximations, the dis-
continuity caused by Tan( / K) being infinite when K = 2
interferes with the solution if the initial estimate of K is on the
wrong side of 2. This engineer derived a simple method of
estimating K which works but lacks the accuracy or simplicity
of the solution in the Engineering Journal.
For those interested, Newtons Method for the sway case
is:
x
i
= x
f(x)
f (x)
= x
[Ax
2
36]Tan (x) 6Bx
[Ax
2
36]Sec
2
(x) + 2AxTan(x) 6B
and for the non-sway case is:
x
i
= x
[Ax
3
2Bx
2
Cot(x) + 2Bx 4x + 8Tan(x / 2
3Ax
2
4BxCot(x) + 2Bx
2
Csc
2
(x) + Tan
2
(x / 2)
where
x
i
= Improved value of x = / K
x = Estmated value of x = / K
A = Gt Gb = (Stiffness Top) (Stiffness Bottom)
B = Gt + Gb = (Stiffness Top) + (Stiffness Bottom)
The iteration will close to any reasonable degree of accu-
racy very rapidly.
CLOSURE BY PIERRE DUMONTEIL
The author wishes to thank Mr. Moore for pointing out that,
should a mathematically exact solution be required, it can be
calculated by Newtons method. The author has found that
using the approximate K factor as a starting value gives a very
rapid convergence.
The approximate formulae are obviously more convenient
for spreadsheet and programmable calculator use.
DISCUSSION
Simple Equations for Effective Length Factors
Paper by PIERRE DUMONTEIL
(3rd Quarter, 1992)
Discussion by William E. Moore II
FIRST QUARTER / 1993 37
CORRECTIONS
Simple Equations for Effective Length Factors
Paper by PIERRE DUMONTEIL
(AISC Engineering Journal, Third Quarter, 1992)
p. III, Equation 4 should read:
K =
3G
A
+ 1.4
3G
A
+ 2.0
p. 112, Table 2 should read:
Table 2.
Comparison of Eqs. 7 and 8Unbraced frames
GA
GB
K exact
K approx.
Error, percent
0.100
0.400
1.083
1.090
1.000
0.250
0.250
1.083
1.100
1.200
0.100
0.900
1.159
1.170
1.000
0.250
0.750
1.162
1.180
1.400
0.500
0.500
1.164
1.180
1.700
0.100
1.900
1.286
1.290
0.300
0.250
1.750
1.295
1.310
0.800
0.500
1.500
1.307
1.330
1.400
GA
GB
K exact
K approx.
Error, percent
1.000
1.000
1.317
1.340
1.800
0.500
4.500
1.575
1.580
0.200
1.000
4.000
1.634
1.650
0.800
2.500
2.500
1.711
1.730
1.200
0.500
9.500
1.777
1.770
0.200
1.000
9.000
1.874
1.880
0.400
2.500
7.500
2.092
2.100
0.600
5.000
5.000
2.228
2.240
0.400
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 38
p. 125: Second Column, Line 12 should read:
d
ht
= diameter of hole in tension plane
(bolt diameter +
1

8
-in.)
p. 125: Second Column, Line 13 should read:
d
hs
= diameter of hole in shear plane
(bolt diamter +
1

16
-in.)
p. 125: Second Column, Line 17 should read:
l
h
= N(1.8 0.6d
hs
) + 0.3d
ns
+ 0.9d + 0.5d
ht
1.8
p. 126: Figure 2 should read:
CORRECTIONS
Fast Check for Block Shear
Paper by LEWIS B. BURGETT
(4th Quarter 1992)
Figure 2
FIRST QUARTER / 1993 39
Eng ine e ring
J ourna l
AMERICAN INSTITUTE OF STEEL CONSTRUCTION, INC.
Pa g e 41: Thoma s Sp uto
De sig n of Pip e Column Ba se Pla te s Und e r Gra vity
Loa d
Pa g e 44: W. Sa mue l Ea ste rling , Da vid R. Gib b ing s,
a nd Thoma s M. Murra y
Stre ng th of She a r Stud s in Ste e l De c k on Comp osite
Be a ms a nd J oists
Pa g e 56: J a me s W. Ma rsh
Ea rthq ua ke s: Ste e l Struc ture s Pe rfor ma nc e a nd
De sig n Cod e De ve lop me nts
Pa g e 66: Ame ric a n Institute of Ste e l Construc tion, Inc .
SI Units for Struc tura l Ste e l De sig n
Pa g e 68: Ne il We xle r
Comp osite Gird e rs with Pa rtia l Re stra ints: A Ne w
Ap p roa c h
2nd Qua rte r 1993/ Volume 30, No. 2
INTRODUCTION
Round pipe columns are often used in large, low buildings
such as warehouses and department stores. No guidance
regarding the design of the base plates for these columns is to
be found in the Manual of Steel Construction,
1
and no defini-
tive guidance may be found in the literature. In lieu of having
a rational method for sizing the base plate, designers have
often resorted to rules of thumb to determine the plate thick-
ness. This paper provides a design procedure for determining
the thickness of these base plates under gravity loads, applied
to Allowable Stress Design criteria. This method is not appli-
cable for uplift conditions where the column is in net tension
nor does it consider erection criteria which is within the
judgement of the detailer and erector.
EFFECTIVE PLATE AREA
For a square base plate under a round column, it is reasonable
to assume the effective bearing area for this analysis to be
within an inscribed circle of the same diameter as the plate
dimensions as shown in Figure 1. Neglecting the area outside
the inscribed circle will have little effect on the design. If a
round base plate is used, the entire area may be considered
effective in the design. In any case, the value of D should be
limited to no greater than 2R.
Therefore:
f
p
= P / D
2
BASE PLATE DESIGN
The base plate can be visualized as being overstressed in two
areas, inside the column diameter and outside the column
diameter. These areas may be designed by applying a yield
line analysis to the base plate.
Yield line analysis of situations similar to this can be found
in many references on yield lines and plastic plate analysis.
2
The simplest method of applying this analysis is that of virtual
work.
Considering first the case of bending within the area en-
closed by the column (Figure 2):
W
i
= Work of the perimeter yielding + work of the radial
lines yielding
= 2M + 2M = 4M
W
e = Bearing pressure volume of cone shown in Figure 2
= f
p
(
1

3
R
2
)
Setting internal work equal to external work:
4M = f
p
(R
2
/ 3)
Allowing M to be equal to the elastic moment capacity of
a rectangle, to be consistent with the 9th Edition method:
M = t
2
F
b
/ 6
setting F
b
= 0.75F
y
and solving for the plate thickness
provides:
t = R(2f
p
/ 3F
y
)
1
2
(1)
Now consider the case of bending of the base plate outside
the column as shown in Figure 3.
W
i
= Work of the radial lines yielding + work of the
perimeter yielding
= 2M[(D R) / D] + 2M(R / D)
= 2M
Thomas Sputo is consulting structural engineer, Gainesville,
FL.
Design of Pipe Column Base Plates
Under Gravity Load
THOMAS SPUTO
Figure 1
SECOND QUARTER / 1993 41
W
e = Bearing pressure volume under the deflected area
outside the column perimeter as shown in Figure 3.
= f
p
[D
2
D
2
/ 3 (D
2
R
2
)(R / D) +
(R
3
/ 3)(R / D)]
= f
p
[2D
2
/ 3 RD + R
3
/ 3D]
Setting internal work equal to external work and inserting
M = t
2
F
b
/ 6 yields:
t = [4f
p
/ 3F
y
(2D
2
3RD + R
3
/ D)]
1
2
(2)
LIGHTLY LOADED BASE PLATES
Reference 3 recommends that the loaded H section method
described in the Manual of Steel Construction be applied to
the design of these base plates. Referring to Figure 4, and
again applying a yield line analysis to the plate:
A
= P / F
p
= (R
o
2
R
c
2
) therefore:
R
c =[R
o
2
(P / F
p
)]
1
2
Applying a yield line analysis to the base plate,
W
i = 2M + 2M = 4M, as for the first case
Figure 2
Figure 3
Figure 4
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 42
W
e
= Bearing pressure under loaded portion of cone in
Figure 4
= F
p
[R
2
/ 3 R
c
2
((R R
c
) / R) ( / 3)R
c
2
(R
c
/ R)]
= F
p
[R
2
/ 3 R
c
2
+ 2R
c
2
/ 3R]
Setting internal work equal to external work, and substitut-
ing for the value of M and F
b
:
t = [(2F
p
/ 3F
y
)(R
2
3R
c
2
+ 2R
c
2
/ R)]
1
2
(3)
It should be noted that if F
p
= f
p
and R
c
= 0, Equation 3
reduces to Equation 1.
If the result of Equation 3 is a plate thickness greater than
that given by Equation 1, Equation 3 does not apply. The
thickness of a lightly loaded base plate should not be greater
than the thickness of one loaded over its entire area.
4
The required base plate thickness is the greater of Equa-
tions 1 or 3 and 2.
EXAMPLE
Given:
Pipe Column, Standard Weight, 4 in. nominal diameter
(OD = 4.500 in., ID = 4.026 in.)
Load = 12 kips
Base Plate = 7 in. 7 in.
Steel = A36
Concrete = 3,000 psi
Solution:
R = (4.500 + 4.026) / 4 = 2.13 in.
D = 7.0 / 2 = 3.5 in.
f
p = 12 kips / (3.5)
2
= 0.31 ksi < 0.35f
c

t = (2.13)[2(0.31) / (3)(36)]
1/2
= 0.161 in. (1)
t
= [(4 / 3)(0.31 / 36)(2(3.5)
2
3(2.13)(3.5) +
(2.13)
2
/ 3.15] = 0.237 in. (2)
F
p = 0.35f
c
= 0 35(3) = 1.05 ksi
R
c = [2.25
2
(12 / (1.05))]
1
2
= 1.19 in.
t
= {[2(1.05) / 3(36)][2.13
2
3(1.19)
2
+
2(1.19)
3
/ 2.13]}
1/2
= 0.191 in. (3)
This column requires a base plate 0.237 inches thick. Use
a plate 77
1

4
-in.
REFERENCES
1. American Institute of Steel Construction, Manual of Steel
Construction, 9th Edition, 1989, Chicago, Ill.
2. Save, M. A. and C. E. Massonnet, Plastic Analysis and
Design of Plates, Shells, and Disks, 1967, Elsevier, New
York.
3. DeWolf, John T., and David T. Ricker, Column Base Plates,
1990, AISC, Chicago, Ill.
4. Ahmed, Salahuddin, and Robert R. Kreps, Inconsisten-
cies in Column Base Plate Design in the New AISC ASD
Manual, Engineering Journal, 3rd Qtr. 1990.
NOMENCLATURE
B = Base plate width
D
= Radius of loaded area of base plate (D 2R)
F
b
= Allowable bending stress in base plate = 0.75F
y
F
p
= Allowable concrete bearing pressure as defined in
manual
f
p
= Actual bearing pressure as defined in manual
M = Base plate internal resisting moment per unit length
P = Total column load
R = Average radius of pipe column
= (R
i
+ R
o
) / 2
R
c
= Inside radius of loaded area for lightly loaded column
R
i
= Pipe column outside radius
R
o
= Pipe column inside radius
t = Thickness of base plate
W
e
= External work done by bearing
W
i
= Internal work done by plate bending
SECOND QUARTER / 1993 43
INTRODUCTION
Composite beam or joist and slab systems typically provide
the most efficient design alternative in steel frame construc-
tion, and indeed it is one of these systems that make steel an
economically attractive alternative to concrete framed struc-
tures. Composite beam specification requirements and design
aids are given in the American Institute of Steel Construction
(AISC) Load and Resistance Factor Design (LRFD) Manual.
1
The LRFD composite beam design procedure results in de-
signs that are typically 1015 percent more economical than
those obtained using the AISC allowable stress design (ASD)
procedure. The efficiency of composite beam design using
LRFD procedures has, in the authors opinions, been the
primary motivating factor for the use of the LRFD specifica-
tion
2
to date.
The design strength and stiffness of composite beams
depends on the shear connection behavior. The strength of the
shear connectors may be reduced because of the influence of
the steel deck geometry. An empirical expression for this
reduction was developed by evaluating results of composite
beam tests in which the deck ribs were oriented perpendicular
to the steel beam.
3
A reduced stud strength is obtained by
multiplying the stud reduction factor, SRF, by the nominal
strength of a shear stud, Q
n
. The expression for the nominal
stud strength,
4
which has been incorporated in the AISC
LRFD specification and is the basis for the tabular values
given in the AISC ASD specification,
5
is given by:
Q
n
0.5A
sc
f
c
E
c
A
sc
F
u
(1)
where
A
sc
= cross-sectional area of a stud shear connector
f
c
= specified compressive strength of concrete
E
c
= modulus of elasticity of concrete
F
u
= minimum specified tensile stress of the stud shear
connector
This equation was developed based on results from elemental
push-out tests.
4
The stud reduction factor is given by:
SRF
0.85
N
r

w
r
h
r
_

H
s
h
r
1.0
_

,
1.0 (2)
where
N
r
= number of studs in one rib at a beam intersection
w
r
= average width of concrete rib
h
r
= nominal rib height
H
s
= length of shear stud after welding
This reduction factor applies to cases in which the deck ribs
are perpendicular to the steel beam and is used in both the
AISC LRFD and ASD specifications.
These equations, or similar forms, have been used in sev-
eral design specifications, both in the United States and
abroad. However, in recent years several researchers
611
have
shown that Equation 2 is unconservative for certain configu-
rations. The studies have considered numerous parameters,
including depth of steel deck shear stud height, concrete unit
weight, position of shear stud in the deck rib relative to the
bottom flange stiffener, number of shear studs in a given deck
rib, and the amount and position of reinforcement in the slab.
The studies reported results from push-out tests alone
6,10,11
or
a combination of push-out tests and beam tests.
79
A conclu-
sion common to all of the studies is that a modified, or
completely different, stud reduction factor is needed. Modi-
fied calculation procedures have been developed and reported
in the recent research studies. However, none of the studies
have reported reasons for the discrepancy between the experi-
mental data and Equations 1 and 2.
The reason for the discrepancy between recent experimen-
tal results with those predicted using Equations 1 and 2 is not
clear. However, it is clear that a significant base of data exists
to substantiate the procedures.
3,12,13
A proper resolution of this
dilemma will require careful consideration of all the data.
A review of the data reported by Grant, et al.,
3
along with
related studies conducted by Henderson
12
and Klyce
13
reveal
W. Samuel Easterling is assistant professor in the Charles E.
Via, Jr. Department of Civil Engineering, Virginia Polytechnic
Institute and State University, Blacksburg, VA.
David R. Gibbings is graduate research assistant in the
Charles E. Via, Jr. Department of Civil Engineering, Virginia
Polytechnic Institute and State University, Blacksburg, VA.
Thomas M. Murray is Montague-Betts professor of structural
steel design in the Charles E. Via, Jr. Department of Civil
Engineering, Virginia Polytechnic, Blacksburg, VA.
Strength of Shear Studs in Steel Deck on
Composite Beams and Joists
W. SAMUEL EASTERLING, DAVID R. GIBBINGS and THOMAS M. MURRAY
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 44
two important characteristics that relate directly to the dis-
crepancy. The majority, but not all, of the tests reported by
Grant, et al. and all the tests reported by Henderson were
detailed such that the studs were placed in pairs within a given
rib. The single test reported by Klyce had two-thirds of the
studs placed in pairs. Also, the deck used in the studies
reported by Grant, et al. did not have a stiffener in the bottom
flange. Both of these details make the position of the shear
stud relative to the stiffener in the bottom flange of the deck,
which is described in greater detail in the following para-
graph, of less concern.
One of the important parameters identified in some of the
recent studies was the position of the shear stud relative to the
stiffener in the bottom flange of the deck. Most deck profiles
manufactured in the United States have a stiffener in the
middle of the bottom flange, thus making it necessary to weld
shear studs off center. Tests have shown differences in shear
stud strengths for the two choices. A stud placed on the side
of the stiffener nearest the end of the span is in the strong
position and one placed on the side of the stiffener nearest the
location of maximum moment is in the weak position. A
schematic of both strong and weak position stud locations is
shown in Figure 1. The difference in strength is partly attrib-
utable to the differences in the amount of concrete between
the stud and the web of the deck that is nearest to mid-span
for the two positions. This detail will be considered further in
subsequent sections of this paper.
A characteristic of partial composite beam design must be
kept in mind when one evaluates results of beam tests and
push-out tests. The relationship between the percentage of
shear connection and the moment capacity is shown in Fig-
ure 2 for a W1631 A36 section. The curves shown in Figure 2
were developed using the calculation procedure in the Com-
mentary to the LRFD specification.
2
The nominal moment
capacity, M
n
, is shown normalized with respect to the fully
composite moment, M
fc
. The percent shear connection is
given by Q
n
/ A
s
F
y
, where Q
n
is the sum of the shear
connector strength between the points of maximum and zero
moment, A
s
is area of steel cross section, and F
y
is yield stress
of the steel cross section. Curves are shown for three values
of Y2, which is the distance from the top of the steel section
to the center of the effective concrete flange. Although the
curves were generated for a W1631, they are representative
of a wide range of cross sections because of the normalization
procedure. A value of M
n
/ M
fc
of about 0.9 is obtained from
a partial shear connection value of 0.7. This relation can be
extended to evaluating test results, in that if a measured to
predicted moment capacity of 0.9 is obtained, then the meas-
ured to predicted shear connector capacity is 0.7. Because of
this relationship, one can argue that an accurate evaluation of
the shear connector strength must be made using carefully
controlled elemental push-out tests, as opposed to evaluating
stud strengths using only beam tests. The sensitivity of the
stud strength to various parameters is difficult to discern if the
strength is back calculated from beam test results. The best
approach is to use a combination of the two test configura-
tions, with the push-out tests being used to evaluate a wide
range of parameters and formulate strength relationships, and
with the beam tests used as confirmatory tests.
The remaining sections of this paper describe a research
project conducted at Virginia Tech to evaluate the strong vs.
weak shear stud position issue.
14
Results from a series of four
composite beam tests are presented. Additionally, the results
from a series of push-out tests are described. The push-out
tests were part of another research project conducted prior to
the beam tests.
15
An analysis of the results is presented which
compares the experimental beam strengths with calculated
values based on Equations 1 and 2, as well as values based on
the push-out tests.
Fig. 1. Strong and weak position shear stud locations. Fig. 2. Normalized moment versus percent shear connection.
SECOND QUARTER / 1993 45
STRENGTH AND STIFFNESS CALCULATION
PROCEDURES
Test results were compared to calculated strength and stiff-
ness values. The calculated shear stud strengths were deter-
mined using the LRFD Specification Equations I5-1 and I3-1
(Equations 1 and 2 in this paper). The flexural strength
calculations were made using the equations given in the
Commentary to the LRFD Specification. The elastic stiffness
values were calculated using the lower bound moment of
inertia defined in Part 4 of the LRFD Manual. Measured
material properties were used in all calculations. The steel
section properties that were measured (depth, flange thick-
ness, flange width, and web thickness) were nearly identical
to the tabular values given in Part 1 of the LRFD Manual.
Therefore, tabulated cross-section properties for the steel
shape were used in the calculations.
The flexural strength calculation procedure gives three
equations for the nominal moment capacity, with the govern-
ing one determined based on the location of the plastic neutral
axis (PNA). Yield stresses were determined separately for the
web and flanges, thus the hybrid section idealization was
used. All the specimens in this study were designed approxi-
mately 40 percent composite and the PNA was located in the
web for all tests. The calculated moment capacity, M
c
, using
Equation C-I3-5,
2
is given by:
M
n
M
p

C
P
yw

_

,
2
M
pw
+ Ce (3)
where
M
p
= steel section plastic moment
C = compressive force in the concrete slab
P
yw
= web yield force
M
pw
= web plastic moment
e = distance from center of steel section to the center of
the compressive stress block in the slab
The force C is given by:
C
min

A
sw
F
yw
+ 2A
sf
F
yf
0.85f
c
A
c
Q
n
(4)
where
A
sw
= area of steel web
F
yw
= yield stress of web steel
A
sf
= area of steel flange
F
yf
= yield stress of flange steel
A
c
= area of concrete slab within effective width
The distance e is given by:
e 0.5d + h
r
+ t
c
0.5a (5)
where
d = depth of steel section
t
c
= slab thickness above the steel deck
a = depth of compression stress block
The lower bound moment of inertia was calculated using the
moment of inertia of the steel beam plus an equivalent area
of concrete, which is a function of the quantity of shear
connection provided. The lower bound moment of inertia,
I
LB
, is given by
I
LB
I
x
+ A
s

Y
ENA

d
2
_

,
2
+

Q
n
F
y

_

,
(d + Y2 + Y
ENA
)
2
(6)
where
I
x
= moment of inertia about x-axis of structural steel
section
Y
ENA
= the distance from bottom of beam to elastic neutral
axis (ENA) and is given by:
Y
ENA

A
s
d
2
+

Q
n
F
y

_

,
(d + Y2)
1
1
]

A
s
+

Q
n
F
y

_

,
1
1
]
(7)
TEST PROGRAM
Beam Test Specimens
The four composite beam tests were similarly constructed.
Each specimen consisted of a single W1631 A36 section
with a composite slab attached. The span of each specimen
was 30 ft and the total beam length was 32 ft because of a 1
ft cantilever at each end. The composite slab used for the
beam tests was constructed using a 20 gage (0.036 in.), 3 in.
deep, composite deck with a total of 6 in. of normal weight
(145 pcf) concrete. The steel deck profile is shown in
Figure 3. A single layer of welded wire fabric (WWF 66-
W1.4W1.4) was placed directly on the top of the deck. A
total of 12 headed shear studs,
3

4
-in. 5 in. after welding, was
used in each test. The studs were welded directly through the
steel deck. The deck was placed with the ribs perpendicular
to the beam span and the slab width was 81 in. A self-drilling
screw was placed in each rib that did not have a shear stud in
it, thus satisfying the requirement of having one fastener
every 12 inches.
16
Deck seams were crimped (button-
punched) twice on either side of centerline, resulting in an
approximately 14-in. spacing. The only nominal difference in
the specimens was the position of the shear studs. However,
the material properties varied for each test.
All of the studs were placed in the strong position for Test 1
and the weak position for Test 2. In Tests 3 and 4 the stud
positions were alternated, thus there were 3 in the strong
position and 3 in the weak position along each half span. The
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 46
stud nearest the support was placed in the strong position and
the stud placement was alternated toward midspan. This
resulted in a symmetric stud pattern in the two half-spans.
(Test 4 was a repeat of the configuration used in Test 3 and
was conducted due to the low concrete strengths obtained in
Test 3.) The ribs in which shear studs were placed are shown
in Figure 4. Note that all of the studs appear in the center of
the deck ribs in Figure 4, however the studs were placed as
described above.
The concrete slabs were formed using 6-in. cold-formed
pour-stop material, resulting in three inches of cover on the
3-in. steel deck. A detail of the deck and slab is shown in
Figure 5. After the concrete was placed, the slab was covered
with plastic and cured for seven days. During this curing time
the slab was kept moist. After seven days, the plastic and the
pour-stop on the sides of the specimen were removed and the
slab was allowed to cure for at least 21 additional days prior
to testing. Concrete cylinders (4 in. 8 in.) were cast at the
same time as the concrete slab. The cylinders were kept
adjacent to the slab, thus were covered with plastic and kept
moist for the initial seven days.
Each specimen was partially supported during construc-
tion. Timber supports were used to prop the steel deck along
the sides of the slab at the quarter points during concrete
placement. This bracing prevented the slab from warping
during the placement of the concrete and was not intended to
shore the beam. The timber props were cut to allow for the
deflection of the beam under the weight of the fresh concrete
and were removed along with the pour-stop after seven days.
Additional support was provided by concrete blocks placed
under the four corners of each specimen to prevent rocking
of the slab during construction and testing.
Beam Instrumentation
A standard instrumentation arrangement for strain, deflec-
tion, end rotation and slip measurement was used for all beam
tests. All of the instruments were monitored using a computer
controlled data acquisition system.
Eight strain gages were used to measure the strain through
the beam cross-section at three different locations, resulting
in a total of 24 strain gages per specimen. Two gages were
placed at each of the following locations: the bottom of the
top flange, the center of the web, the top of the bottom flange
and the bottom of the bottom flange, as indicated in Figure 6.
Gages were placed near one end support, at one quarter point
and at the centerline.
Vertical deflections were measured at the centerline and the
quarter points. Measurements were taken using linear wire
transducers.
Slip measurements were made using potentiometers at-
tached to the top flange of the beam. The potentiometers
measured the relative movement between the top flange of
the beam and a screw embedded in the concrete slab through
a hole in the steel deck. A total of 12 potentiometers were used
in each test, except Test 1, with one placed adjacent to each
shear stud. Slip was not measured adjacent to the two studs
nearest to midspan in Test 1. The slip measurement detail is
shown in Figure 7.
End rotations were measured using two different tech-
niques. Transducers were used to measure the upward deflec-
tion of the ends of the specimen and the support beam. The
1 ft overhang was assumed to rotate rigidly about the support,
Fig. 3. Composite deck profile.
Fig. 4. Shear stud locations for composite beam specimens.
Fig. 5. Deck/slab detail.
Fig. 6. Strain gage locations for composite beam specimens.
SECOND QUARTER / 1993 47
thus using the net upward deflection and the distance between
the measurement and the support, the end rotation was calcu-
lated. Additionally, a digital level was used to measure the
angle of the slab relative to horizontal, over the support, to
the nearest 0.1 degrees.
In addition to the strain measurements already described,
axial strain was measured in a select number of studs in
Tests 24. This measurement was made using an innovative
approach, adapted from bolt strain measurement techniques.
However, due to problems with the gage installation tech-
nique, only a limited amount of usable data was obtained. For
the benefit of those involved with similar research in the
future, the instrumentation technique is presented here.
A cylindrical uniaxial strain gage, referred to as a bolt gage
by the manufacturer, was inserted in the stud into a pre-drilled
hole (approximately 0.1-in. diameter) after it had been welded
to the beam. Lead wires were attached and electrical shrink
tubing was placed over the lead wires to protect them during
concrete placement. The end of the shrink tubing was embed-
ded in a small amount of protective coating that was applied
to the top of the stud. Subsequently the tubing was heated to
conform to the general shape of the lead wire bundle. The lead
wires were brought from the gage straight up through the
concrete to prevent interference with the bonding between the
concrete and the shear stud. A detail of the strain-gaged shear
stud is shown in Figure 8.
The problems with the installation technique were attrib-
uted to the method used to insert the glue in the pre-drilled
hole. The viscosity of the glue was such that the glue had to
be worked into the hole using a blunt probe. Once the gage
was inserted, it was worked back and forth to eliminate any
air bubbles. A different technique, which utilizes a syringe to
fill the hole from the bottom, has been used in other tests on
composite members since the completion of the beam tests.
The change in installation procedures appears to have cor-
rected the problem.
Beam Load Apparatus and Test Procedure
A four-point loading system was used for all tests, with the
loads spaced seven feet apart. The load was applied with a
single hydraulic ram and distributed to the slab by a two-tier
distribution system, as shown in Figure 9.
The load program was similar for all tests. An initial load,
equal to approximately 15 percent of the calculated strength,
was applied to seat the specimen and was then removed. The
instrumentation was then re-initialized. Load increments
were applied to the specimen until the load vs. centerline
displacement response became non-linear. The specimen was
then unloaded and then reloaded to the previous peak in three,
approximately equal, increments. Displacement increments,
based on the mid-span vertical deflection, were subsequently
used to complete the test. The specimen was unloaded during
Fig. 7. Slip measurement detail.
Fig. 8. Detail of strain gage in a shear stud. Fig. 9. Loading frame for composite beam specimens.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 48
the displacement controlled phase if it was necessary to adjust
the loading apparatus.
Push-Out Test Specimens
A total of eight push-out specimens were fabricated, four with
studs in the weak position and four with studs in the strong
position. These tests were performed as part of another study
reported by Sublett, et al.
15
The push-out tests were con-
structed using the same deck profile and shear stud size that
were used in the beam tests. Each half of a push-out specimen
was constructed by attaching a piece of 3-in. deep composite
steel deck to a W511. The ribs of the deck were oriented
perpendicular to the length of the WT section. One or two
shear studs (
3

4
-in. 5 in. after welding) were welded through
the deck to the structural tee. Two each of the strong and weak
position groups had one stud per specimen half. The other two
specimens in each group had two studs, spaced 12 inches
apart along the length of the WT, on each specimen half. A
normal weight concrete slab, 6-in. thick by 24-in. wide by
36-in., was cast on the deck. Welded wire fabric (WWF
66-W1.4W1.4) was placed on top of the deck prior to
casting the concrete. The specimens were covered and kept
moist for seven days, at which time the forms were removed.
Concrete test cylinders (4 in. 8 in.) were cast along with the
push-out specimens and cured in a similar manner.
After the slabs had cured, two halves were bolted through
the stems of the structural tees to form a complete specimen.
This manner of casting permitted the slabs to be cast horizon-
tally and from the same batch of concrete. By doing this the
concrete curing problems associated with either casting the
specimens vertically or from different mixes were avoided.
Overlapping the stems of the tees induced an eccentricity in
the built-up steel section, as compared to using a rolled
H-shape. The effect due to this eccentricity was deemed
negligible.
Push-Out Test Instrumentation
A standard instrumentation arrangement for measurement of
slip, shear load, and normal load was used for all tests. Slip
between the steel deck and steel section was measured at two
locations on each half of the push-out specimen using me-
chanical dial gages. The applied shear load was measured
using a load cell that was part of the universal test machine.
A normal force was applied to the slab, as described in the
next section of the paper, and monitored using a electronic
load cell.
Push-Out Load Apparatus and Test Procedure
To prevent premature separation between the slab and steel
deck, in a direction normal to the slab surface, a yoke device
was placed on the specimen. This manner of loading simu-
lated the gravity load placed on a slab in a composite
beam/slab arrangement. A load cell and hydraulic ram were
part of the yoke assembly. The specimen configuration with
the yoke in place is shown in Figure 10.
Specimens were placed in a universal testing machine on
an elastomeric bearing pad, which minimized the effects
caused by any unevenness in the bottom of the specimen.
Shear load was applied with the universal testing machine in
load increments equal to approximately 10 percent of the
expected specimen capacity. Displacement control was used
once the load levels reached approximately 80 percent of the
expected capacity.
Load normal to the slab surface was applied using the yoke
assembly. The load was monitored using a load cell and
controlled with a hydraulic hand pump and ram. The normal
load was increased along with the applied shear load. The
normal load was approximately 10 percent of the applied
shear load throughout a test.
Material Tests
Standard material tests were conducted on the concrete and
steel components. The concrete cylinders were tested to de-
termine compressive strength on the days of the various beam
and push-out tests. Tensile coupons (0.5 in. width, 2 in. gage)
were cut and machined from both the web and one flange of
each structural steel shape, as well as from flat widths of the
steel deck profile. The ultimate tensile stress for the shear
Fig. 10. Push-out specimen schematic.
SECOND QUARTER / 1993 49
studs was reported by the manufacturer. Material properties
are given in Table 1.
TEST RESULTS
Beam Test Results
The observed behavior was similar for all beam tests, but
notable differences exist. A normalized moment versus de-
flection plot of the four tests is shown in Figure 11. The
experimental moments, M
e
, were normalized with calculated
moment strengths using measured material properties and the
procedure described previously. Note that the plots in Fig-
ure 11 include the non-composite load and corresponding
deflection. The vertical mid-span deflection, , was normal-
ized with
H
. The deflection corresponding to the point where
the elastic stiffness, calculated using the lower bound moment
of inertia, intersects the calculated moment strength is defined
as
H
.
As indicated in Figure 11, all tests exhibited a ductile
response. The moment versus deflection response in Tests 1, 3,
and 4 (strong and alternating stud position tests) remained
elastic up to a normalized moment of approximately 0.6.
Test 2 (weak stud position test) remained elastic up to a
normalized moment of approximately 0.4.
The behavior of the shear studs was distinctly different for
the strong and weak position studs. Strong position studs
exhibited failure by developing concrete shear cones or by
shearing off in the shank. Weak position studs exhibited
failure by punching through the deck rib without developing
a significant shear cone in the concrete or shearing in the stud
shank. In Tests 1, 3 and 4, one or two of the strong position
studs closest to one of the specimen supports sheared off in
the shank. However, the weak position stud between the two
strong position studs in Tests 3 and 4 did not shear off, but
punched through the deck web and remained attached to the
beam.
Push-Out Test Results
An average strength of 13.55 kips per stud was obtained from
the four push-out tests in which the studs were in the weak
position. The concrete compressive strength was similar for
each of the tests, with an average for the four tests of 4.27 ksi.
There was no significant difference between the strengths
(load per stud) obtained from the tests with one stud per
specimen half and the tests with two studs per specimen half.
In all of the weak position tests, failure occurred by the studs
punching through the adjacent web of the steel deck. A small
wedge of concrete between the stud and the deck web was
crushed or broken out in each of the tests. The deck was
noticeably bulged out adjacent to the stud prior to reaching
the maximum applied shear load. This behavior was an indi-
cation that the load was being primarily resisted by the deck.
An average of 18.82 kips per stud was obtained from the
three push-out tests in which the studs were in the strong
position. The average concrete strength was 4.57 ksi. The
results for the fourth specimen were inexplicably low and are
not included in the average. The decision to omit this test was
based on the other three tests plus an additional 11 tests,
similarly constructed, that were part of a proprietary study in
which double angle sections were used as the base members
instead of structural tees. There was no significant difference
between the strengths (load per stud) obtained from the test
with one stud per specimen half and the tests with two studs
per specimen half. In all of the strong position tests, the
strength was limited by the development of a failure surface
Fig. 11. Normalized midspan moment versus
displacement for composite beam specimens.
Table 1.
Material Properties for Composite Beam Specimens
Test
Flange
F
yf
(ksi)
Flange
F
uf
(ksi)
Web
F
yw
(ksi)
Web
F
uw
(ksi)
Slab
f
c

(ksi)
1
2
3
4
42.0
41.9
42.5
43.6
68.8
70.4
70.1
63.4
47.0
45.4
47.0
49.1
71.9
73.8
75.7
62.9
4.81
3.20
2.28
4.99
Shear Studs: Fu = 64.8 ksi
Steel Deck: Fy = 40.3 ksi Fu = 53.6 ksi
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 50
in the concrete. None of the shear studs exhibited a shear
failure in the shank.
The response of the studs in the weak position, in terms of
load versus slip, was more ductile than that of the studs in the
strong position. This difference is attributed to the way in
which the load appeared to be resisted, based on the observed
failure modes. The failure mode for the strong position tests
was brittle; concrete shear, and the failure mode for the weak
position tests was more ductile; bearing and eventual tearing
of the steel deck web. A typical plot of load versus slip
behavior for strong and weak position shear studs is illus-
trated in Figure 12.
ANALYSIS OF RESULTS
The results of the beam and push-out tests were compared
with calculated values. Several comparisons have been made
and are presented in this section. The calculated moment
values were based on the expressions described previously in
this paper, using measured material properties and values of
shear connector strength that were calculated using the LRFD
specification or taken from normalized push-out test results.
Shear connector strength was also back calculated using the
experimental moment values obtained from the beam tests.
The results of each of these calculations and comparisons are
given in Table 2.
The values Q
c
given in Table 2 are calculated stud strengths.
These were determined using Equations 1 and 2 with meas-
ured material properties. Stud strengths Q
cb
, were back-
calculated using the experimental moment from the beam
tests, measured material properties and the calculation proce-
dure described previously.
Because the shear studs in the weak position, in both the
push-out and beam tests, failed by punching through the web
of the deck it was hypothesized that their strength was not
primarily a function of concrete strength. Rather, the stud
strength is primarily a function of the steel deck strength (i.e.,
the yield stress of the steel deck). Certainly some interaction
between the concrete and the deck occurred, but the dominant
component was the steel deck. Based on this hypothesis, the
weak position push-out test strengths were averaged and used
for all the weak position stud strengths in the calculations for
the beam tests. No adjustment was made to account for
variable concrete strengths.
The strength of the shear studs in the strong position was
taken as a function of the concrete strength. The strong
Table 2.
Experimental and Calculated Results
Test
Q
c
(kips)
Q
po
(kips)
Q
cb
(kips)
M
c
(ft-kips)
M
po
(ft-kips)
M
e
(ft-kips)
Q
cb
/ Q
c
Q
po
/ Q
c
Q
cb
/ Q
po
M
e
/ M
c
M
e
/ M
po
1 (str.) 28.7 19.3 18.8 344 303 304 0.66 0.67 0.97 0.88 1.00
2 (weak) 22.6 13.6 13.4 316 274 273 0.59 0.60 0.99 0.87 1.00
3 (alt.) 17.5 13.3 14.5 297 277 283 0.83 0.76 1.09 0.95 1.02
4 (alt.) 28.7 16.6 17.0 354 301 303 0.59 0.58 1.02 0.86 1.01
All values based on measured material properties
Qc = calculated stud strength using Equations 1 and 2.
Qpo = calculated stud strength using Equation 8 and concrete strength from beam test for strong position studs and a constant value of 13.55 kips
for the weak position studs.
Qcb = calculated stud strength using Equation 3 with Me in place of Mn.
Mc = calculated moment strength using Equation 3 and Qc.
Mpo = calculated moment strength using Equation 3 and Qpo.
Me = maximum applied experimental moment including weight of specimen, load beams, and applied ram load.
Fig. 12. Load vs. slip for strong and weak
position shear studs for push-out tests.
SECOND QUARTER / 1993 51
position stud strengths in the beam tests were calculated by
normalizing the push-out test results with the concrete
strengths as given by:
Q
po
18.82 kips

f
c

4.57 ksi
(8)
where f
c
is the concrete compressive strength for the compos-
ite beam test, 18.82 kips is the average stud strength from the
push-out tests, and 4.57 ksi is the concrete compressive
strength from the push-out tests. The Q
po
values represent stud
strengths for the beam tests based on push-out test results.
Equation 8 was used to calculate the values for Q
po
in the
Test 1, and the constant value reported in the push-out results
section was used for Test 2. The strong and weak position
values were averaged in determining the values for O
po
in
Tests 3 and 4.
Three values of moment are shown in Table 2, M
c
, M
po
, and
M
e
. The first, M
c
, was calculated using Q
c
, M
po
was calculated
using Q
po
, and M
e
represents the maximum experimental
moment from the beam tests. Various ratios of stud strengths
and moment strengths are also given in Table 2.
Two trends are clearly indicated by the results in Table 2.
One of these is that the stud strengths predicted by Equa-
tions 1 and 2 do not compare favorably to the values from the
push-out tests or the beam tests. This is indicated by the ratios
Q
cb
/ Q
c
and Q
po
/ Q
c
. The second trend that is evident is that
the results from the push-out tests and beam tests compare
very well, as indicated by the ratio Q
cb
/ Q
po
.
Additionally, while a comparison between strong and weak
position shear stud strengths indicates some difference, the
more pronounced and significant difference is between the
predicted values and the beam and push-out test results. The
ratios Q
cb
/ Q
c
or Q
po
/ Q
c
indicate the strong position values
are approximately 70 percent of the predicted and the weak
position values are approximately 60 percent of predicted.
The sensitivity of the moment strength to the shear stud
strength is also illustrated in the results. Values of experimen-
tal to calculated shear stud strengths varied between 0.59 and
0.83, while the experimental to calculated moment values,
indicated by M
e
/ M
c
, varied between 0.85 and 0.94. The
relationship between shear connection and moment strength
is illustrated for the W1631 used in this study by the nor-
malized moment versus shear connection relationship in Fig-
ure 2. Although as previously indicated, this relationship is
generally presented in the context of partial composite design,
it can also be used to consider the reduction in moment
strength due to a reduction in shear connector strength.
The strain data collected from the beam tests also indicate
the difference between strong, weak, and alternating position
studs. The relationship between the position of neutral axis
and the applied moment is illustrated in Figure 13. A linear
regression analysis was performed using the eight strain
readings located at midspan in the steel section to determine
the neutral axis location. As noted in Figure 13, the strong
position studs resulted in the neutral axis being higher in the
steel than for the weak or alternating tests. Further, the posi-
tion for the alternating tests fell between the strong and weak
values.
Using Figure 13, the plastic neutral axis position can be
established by visually locating the point at which the slope
of line is approximately vertical. These values are given in
Table 3. Also shown in Table 3 are calculated values of the
plastic neutral axis based on Q
c
, Q
po
, and Q
cb
. Note that the
calculated values using either Q
po
or Q
cb
correspond more
closely to the experimental values than do the positions
calculated using Q
c
, in all but Test 3.
DESIGN IMPLICATIONS
The implications of the study described here, as well as
previous studies, on composite beam design merit considera-
Fig. 13. Applied moment versus position of neutral
axis for composite beam specimens.
Table 3.
Experimental and Calculated Neutral Axis Positions
Test
PNA
e
(in.)
PNA
c
(in.)
PNA
po
(in.)
PNA
cb
(in.)
1
2
3
4
2.7
4.2
3.9
3.6
0.78
2.32
3.88
0.88
3.13
4.56
4.64
3.85
3.15
4.61
4.57
3.76
All values of PNA are measured from top of steel section.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 52
tion at this point. Based on the test results presented in the
previous sections, it is evident that Equation 2 is not conser-
vative in all cases. Specifically, if single shear studs are used,
as opposed to pairs of studs, the equation over-predicts the
strength of the stud. Based on a review of previous stud-
ies
3,12,13
the authors believe that Equation 1 is conservative for
designs in which two studs per rib are utilized. No general
modifications to the form of the equation are proposed at this
time. Until such modifications are formulated, the following
recommendations are offered:
1. The stud reduction factor should not exceed 0.75 for
cases in which there is one stud in a rib.
2. Detail all single studs in the strong position. The imple-
mentation of this detail requires coordination between
the structural engineer and the stud contractor to effec-
tively relay the objective of the detail.
3. Use 50 percent composite action as a minimum, i.e.,
keep Q
n
/ A
s
F
y
greater than or equal to 0.50. This will
minimize the adverse effect of under-strength studs on
the design moment strength, as reflected by the trend of
the curves in Figure 2.
The result of implementing the above recommendations is
an increase in the number of shear studs for designs utilizing
one stud per rib. This will obviously result in a small increase
in the cost, however the percentage increase in the in-place
cost of the composite beam for these situations will be minor.
Certainly in view of the questions that have been raised
regarding the strength of the studs, the increase is warranted.
A consideration in future composite beam studies and
modifications to the specification procedures should be the
application of a strength reduction factor, , to the shear studs.
In the current AISC LRFD specification
2
a single strength
reduction factor is applied to the nominal moment strength
for the composite beam system, which includes the variable
effects of the shear connectors. However, the flexural strength
of the beam and the shear strength needed at the steel concrete
interface are associated with different modes of behavior and
limit states and therefore merit separate consideration. If this
approach were pursued, one would expect that the value of
for the flexural limit state may increase above the present
value of 0.85, thus making more efficient use of the steel
shape which is the dominant component in the cost of the
composite bearn. At the same time the variability that exists
in the shear stud strength would be reflected in a value for
shear studs.
The flexural and shear stud limit states are treated inde-
pendently in other limit states design specifications.
17,18
The
nominal strengths, as well as the stud reduction factors, vary
between the three specifications. A graphical comparison of
the three specifications for the 3-in. deep composite deck
shown in Figure 3 is given in Figure 14. The differences
illustrated in Figure 14 in part reflect the uncertainty that
exists at the present time regarding shear connector strength.
SUMMARY AND CONCLUSIONS
Results were described for a recent study conducted at Vir-
ginia Tech in which a series of push-out tests and composite
Fig. 14. Shear strength comparison for AISC, CSA, and Eurocode specifications.
SECOND QUARTER / 1993 53
beam tests were conducted. The results were consistent with
other recent studies reported in the literature, in that the
strength of shear studs placed in the ribs of steel deck oriented
transverse to the beam span, calculated using Equation 2,
were higher than measured values. Review of the test data
used to develop Equation 2 indicated that the majority of the
tests were conducted with the shear studs placed in pairs.
Equation 2, when combined with Equation 1, accurately
reflects the stud strength for these cases.
Specific modifications to Equation 2 were not proposed, as
further evaluation of existing procedures is required. The
hypothesis regarding the influence of the steel deck material
properties on the stud strength must be evaluated at the same
time and perhaps included as a modification to one of the
existing methods. This hypothesis, while not conclusively
verified, was supported by the results of the Virginia Tech
research program.
ACKNOWLEDGMENTS
Graduate research assistant support for the project was pro-
vided by the American Institute of Steel Construction. The
following organizations generously supplied material and
equipment for the project: Virginia-Carolinas Structural Steel
Fabricators Association (structural steel), Vulcraft Division
of Nucor (steel deck pour-stop and welded wire fabric), and
Nelson Stud Welding Division of TRW (shear studs and stud
welding equipment). The remaining project costs were pro-
vided by Virginia Tech. The project from which the push-out
test results were taken was sponsored by Nucor Research and
Development.
REFERENCES
1. American Institute of Steel Construction, Manual of Steel
ConstructionLoad and Resistance Factor Design, First
Edition, Chicago, Illinois, 1986.
2. American Institute of Steel Construction, Load and Re-
sistance Factor Design Specification for Structural Steel
Buildings, Chicago, Illinois, September 1986.
3. Grant, J. A., Fisher, J. W. and Slutter, R. G, Composite
Beams with Formed Steel Deck, Engineering Journal,
AISC, 14(1), 1977, pp. 2443.
4. Ollgaard, J. G., Slutter, R. G. and Fisher, J. W., Shear
Strength of Stud Connectors in Lightweight and Normal
Weight Concrete, Engineering Journal, AISC, 8(2),
1971, pp. 5564.
5. American Institute of Steel Construction, Specifications
for Structural Steel Buildings: Allowable Stress Design
and Plastic Design, Chicago, Illinois, June 1989.
6. Hawkins, N. M. and Mitchell, D., Seismic Response of
Composite Shear Connections, Journal Structural Engi-
neering, ASCE, 110(9),1984, pp. 21202136.
7. Jayes, B. S. and Hosain, M. U., Behaviour of Headed
Studs in Composite Beams: Push-out Tests, Canadian
Journal of Civil Engineering, 15, 1988, pp. 240253.
8. Jayes, B. S. and Hosain, M. U., Behaviour of Headed
Studs in Composite Beams: Full-Size Tests, Canadian
Journal of Civil Engineering, 16, 1989, pp. 712724.
9. Robinson, H., Multiple Stud Shear Connections in Deep
Ribbed Metal Deck, Canadian Journal of Civil Engi-
neering, 15, 1988, pp. 553569.
10. Mottram, J. T. and Johnson, R. P., Push Tests on Studs
Welded Through Profiled Steel Sheeting, The Structural
Engineer, 68(10), (1990), pp. 187193.
11. Lloyd, R. M. and Wright, H. D., Shear Connection
between Composite Slabs and Steel Beams, Journal of
Construction Steel Research, 15, 1990, pp. 255285.
12. Henderson, W. D., Effects of Stud Height on Shear
Connector Strength in Composite Beams with Light-
weight Concrete in Three-Inch Metal Deck, Master of
Science Thesis, The University of Texas, Austin, TX,
1976.
13. Klyce, D. C., Shear Connector Spacing in Composite
Members with Formed Steel Deck, Master of Science
Thesis, Lehigh University, Bethlehem, PA, 1988.
14. Gibbings, D. R., Easterling, W. S. and Murray, T. M.,
Composite Beam Strength as Influenced by the Shear
Stud Position Relative to the Stiffener in the Steel Deck
Bottom Flange, Report No. CE/VPI-ST 92/07. Virginia
Polytechnic Institute and State University, Blacksburg,
VA, 1992.
15. Sublett, C. N., Easterling, W. S. and Murray, T. M.,
Strength of Welded Headed Studs in Ribbed Metal Deck
on Composite Joists, Report No. CE/VPI-ST 92/03, Vir-
ginia Polytechnic Institute and State University,
Blacksburg, VA, 1992.
16. American Society of Civil Engineers, Specifications for
the Design and Construction of Composite Slabs,
ANSI/ASCE 3-84, New York, 1984.
17. Commission of the European Communities, Eurocode 4:
Common Unified Rules for Composite Steel and Concrete
Structures, Rep. EUR 9886, 1992.
18. Canadian Standards Association, Limit States Design of
Steel Structures, CAN/CSA-S16.1-M89, Rexdale, On-
tario, 1989.
NOMENCLATURE
A
c
= area of concrete slab within effective width
A
s
= area of steel cross section
A
sc
= cross sectional area of a stud shear connector
A
sf
= area of steel flange
A
sw
= area of steel web
a = depth of compression stress block
C = compressive force in concrete slab
d = depth of steel section
E
c
= modulus of elasticity of concrete
e = distance from center of steel section to the center of
the compressive stress block in the slab
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 54
F
u
= minimum specified tensile stress of stud shear
connector
F
y
= yield stress of steel cross section
F
yf
= yield stress of steel web
F
yw
= yield stress of steel web
f
c

= specified compressive strength of concrete


H
s
= length of shear stud after welding
h
r
= nominal rib height
I
LB
= lower bound moment of inertia
I
x
= moment of inertia about x-axis of structural steel
section
M
c
= moment strength calculated using Q
c
M
e
= maximum experimental moment
M
fc
= fully composite moment strength
M
n
= nominal moment strength
M
p
= steel section plastic moment strength
M
po
= moment strength calculated using Q
po
M
pw
= web plastic moment
N
r
= number of studs in one rib at a beam intersection
P
yw
= web yield force
Q
c
= calculated stud strength using Equations 1 and 2
Q
cb
= stud strength calculated using M
e
and Equation 3.
Q
po
= stud strength calculated using push-out test results
Q
n
= nominal strength of a shear stud
t
c
= slab thickness above the steel deck
w
r
= average width of concrete rib
Y
con
= distance from top of steel beam to top of concrete
Y
ENA
= distance from bottom of beam to elastic neutral axis
Y2
= Y
con
a / 2
Q
n
= sum of strengths of shear connectors
SECOND QUARTER / 1993 55
INTRODUCTION
Major earthquakes occur several times each year through-
out the world with heavy loss of life and property. Recent
examples are the 1992 Cairo, Egypt earthquake with the loss
of over 500 lives, and the Mexico earthquake of 1985 with
the loss of 8,000 lives and the collapse of over 400 buildings.
The United States has experienced many large earth-
quakes, with the most seismic activity to date being located
in California, i.e., Loma Prieta, California, 1989, 7.1 Richter
magnitude and Landers, California, 1992, 7.5 Richter magni-
tude. It is evident from past occurrences of earthquakes that
the highly seismic regions of the United States have a serious
earthquake problem, and the less serious regions in the central
and eastern parts of the country now realize that they have an
earthquake problem which is being addressed through adop-
tion of the latest seismic design provisions into the BOCA and
SBCCI building codes. Some of these newly acquired seismic
provisions are taken from the Building Seismic Safety Coun-
cil program on improved seismic safety.
The Building Seismic Safety Council (BSSC) was estab-
lished in 1979 under the auspices of the National Institute of
Building Sciences (NIBS) as an entirely new type of instru-
ment to develop and promulgate building earthquake hazard
mitigation regulatory provisions that are national in scope. Its
fundamental purpose is to enhance public safety by providing
a national form that fosters improved seismic safety provi-
sions for use by the building community in the planning,
design and construction of buildings. To fulfill its purpose,
the BSSC promotes the development of seismic safety provi-
sions suitable for use throughout the United States. The BSSC
believes that the regional and local differences in the nature
and magnitude of potentially hazardous earthquake events
require a flexible approach to seismic safety that allows for
consideration of the relative risk, resources and capabilities
of each community. The BSSC itself assumes no standards-
making and promulgating role; rather, it advocates that code
and standards formulation organizations consider BSSC
recommendations for inclusion into their documents and stand-
ards. A recommendation that is taking place today in code writing.
The basic problem of earthquake design is to synthesize the
structural configuration; the size, shape, and material of the
structural elements along with the methods of fabrication, so
that the structure will safely and economically withstand the
action of earthquake ground motions. This of course requires
a broad knowledge of the behavior of structures during earth-
quakes, and the final evaluation of the design will be made
by a future earthquake. It is this ultimate test that has shown
that steel-framed buildings and bridges have an excellent
record of protecting life safety, as well as minimizing eco-
nomic loss and business interruption.
1
STRUCTURES PERFORMANCE
Mexico Earthquake
On September 19, 1985, a magnitude 8.1 earthquake struck
Mexico, followed by a magnitude 7.5 aftershock 36 hours
later.
2
Data compiled by the Institute of Engineering of the
National Autonomous University of Mexico revealed that a
total of 330 buildings collapsed in central Mexico City. Of
these, 12 were steel frame, and 318 were reinforced concrete
or masonry. The majority of steel buildings that collapsed
were built before 1957, while most concrete and masonry
buildings were built between 1957 and 1976. Both years mark
major revisions in the building code, adopted in response to
past damaging earthquakes.
Steel structures constructed after 1957 fared much better
than the norm, with only 6.8 percent of such buildings expe-
riencing severe damage or collapse. Steel structures con-
structed after 1976 performed excellently; no cases of severe
damage or collapse were noted in this group, and only four
such buildings sustained any structural damage.
2
Some steel buildings constructed from 1920 through the
1940s experienced severe damage. These structures were
built prior to the adoption of earthquake codes, and used
construction types that were abandoned following the 1957
earthquake. The most common type of steel building con-
struction used over the last three decades has been highly
redundant moment frames where almost every beam-column
joint in these structures is moment resisting. The second most
common lateral system for steel structures was moment-
resisting frames with braced bays. The Pino Suarez complex
accounted for all the reported failures of this system in the
1985 earthquake.
Analyses performed after the earthquake have provided an
explanation of the Pino Suarez failures.
3
Very large axial
Earthquakes: Steel Structures Performance and
Design Code Developments
JAMES W. MARSH
James W. Marsh is a professional engineer in El Monte, CA.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 56
loads, due to gravity and seismic overturning, overstressed
the exterior columns in the braced bays. The bracing system
being capable of resisting story shears several times higher
than the code design level produced unanticipated large axial
forces in the columns. As the result of these findings, provi-
sions have been added to the 1988 edition of the Uniform
Building Code to prevent overload of columns from overturn-
ing forces that exceed those calculated from the basic seismic
provisions of the Code.
Structural steel was successfully used to strengthen rein-
forced concrete buildings prior to the 1985 earthquake.
4
The
12-story Durango Building is located in the heaviest damaged
region of the city. After being heavily damaged in the 1979
earthquake, the building was retrofitted with steel frames,
which added ductility as well as strength to the structure. In
the 1985 earthquake the building performed excellently, sus-
taining no damage. The steel frames are believed to have
carried over 80 percent of the total lateral force.
Although steel construction in Mexico differs substantially
from practice in the United States, dozens of modern steel
buildings located in the badly shaken lake bed area of Mexico
City received no damage. A good example is the 44-story
Torre Latinoamericana, constructed in the early 1950s and
designed for earthquake loads, which performed excellently
in 1985 as it had in three previous earthquakes in 1957, 1978
and 1979.
Whittier-Narrows, California Earthquake
The October 1, 1987 Whittier-Narrows earthquake of magni-
tude 5.9 (Richter Scale) was considered a moderate earth-
quake. Several aftershocks caused a few structures that were
badly damaged on October 1 to collapse in an October 4
aftershock of 5.5 magnitude. USGS records show unusual
high ground accelerations of 0.40g to 0.60g, and ground
displacements of 1 to 2 inches.
5
According to the National
Center for Earthquake Engineering Research (NCEER), most
earthquake damage occurred in unreinforced masonry build-
ings, older homes and modern buildings in construction types
lacking in ductility.
6
Several reinforced concrete and shear wall buildings,
bridges constructed according to pre-1971 engineering prac-
tice sustained heavy damage. Major shear damage was expe-
rienced by the supporting concrete columns of the overpass
located at the junction of the I-5 and I-605 Freeways, 15 miles
East of downtown Los Angeles. Whereas bridge abutments
experienced moderate to minor damage by spauling of con-
crete underneath the supporting pads, no damage was noted
in abutments, columns and piers of bridges that were retrofit-
ted by cable restrainers.
6
A two-story concrete parking structure built in 1964 and
located in the Whittier Quad shopping center collapsed after
shear failure of its columns. Large girders had much stronger
sections than the supporting columns, thereby creating a
strong-beamweak-column situation. The requirement for a
strong-column vs. a weak-beam design was first required for
concrete structures in the 1985 Uniform Building Code. The
Code requires that the sum of the column moments at a
beam-column joint be a minimum of 20 percent greater than
the sum of the girder moments. A similar design provision
became a requirement for structural steel seismic design in
1988.
7
The four-story steel-framed California Federal Savings
Service Center relied upon braced (chevron) frames for lateral
resistance and was designed in accordance with the 1979
Uniform Building Code. During the earthquake the building
experienced a peak ground motion several times higher than
the working stress design levels, Structural damage was lim-
ited to the buckling of a single wide flange bracing member
on each of the second, third and fourth floors.
8
In spite of the
severe ground motions that the building experienced, it was
restored to service within a week. In contrast, an adjacent
two-story precast concrete structure built in 1980 experienced
such extensive damage that repair to the building took nine
months.
Loma Prieta Earthquake
On October 17, 1989 an earthquake of 7.1 Richter magnitude
occurred that was centered approximately 60 miles south of
San Francisco. Among the seismic-induced events were the
collapse of the elevated Cypress Street section of Interstate
880 in Oakland; the collapse of a section of the San Francisco-
Oakland Bay Bridge; and major structural damage to modern
buildings in Oakland, San Francisco and Burlinghame. Over
62 people died.
Some of the heaviest concentration of damage occurred in
the city of Oakland, 60 miles north of the earthquake epi-
center, where peak ground accelerations were only 0.2g to
0.26g.
9
A 15-story concrete shear wall structure in downtown
Oakland suffered extensive damage when its lightweight
concrete shear walls shattered at the first story. The presence
of a redundant steel frame within the building saved the
structure.
2
The Hyatt Regency Hotel located in Burlingame, a mid-
Table 1.
Statistical Summary of Damage to Buildings
1985 Earthquake
Type
Structure
Extent of
Damage
Year When Built
Pre-1957 5776 Post 1976 Total
Steel
Frame
Collapse
Severe
7
1
3
1
0
0
10
2
RC Frame Collapse
Severe
27
16
51
23
4
6
82
45
Waffle
Slab
Collapse
Severe
8
4
62
22
21
18
91
44
SECOND QUARTER / 1993 57
rise reinforced concrete structure, sustained extensive dam-
age to its shear walls and floor slab around the elevator core.
The structure was designed to the 1985 Uniform Building
Code and construction completed just prior to the earthquake.
Repair of damage resulted in closure of the hotel for more
than eight months. In contrast, modern steel-framed buildings
performed excellently in the Loma Prieta earthquake, as they
have in the past.
Damage to steel structures was typically limited to crack-
ing of cladding and interior partitions with wide-spread dis-
array of building contents. The nonstructural damage sus-
tained by steel frame buildings may largely be attributed to
their flexibility, which results in large displacements.
10
Major transportation routes were affected by the Loma
Prieta earthquake. Immediately after the earthquake 11 major
highways and freeways were closed due to landslides, struc-
tural damage or bridge collapse. The collapse of the Cypress
Street elevated section of I-880 (near downtown Oakland)
was responsible for the majority of earthquake deaths. The
double-deck highway system consists of box girder decks
supported by concrete frames. The failure occurred at the
connection of the support columns and the transverse beams,
at the lower road level.
Redesign of the Cypress Street roadway was completed in
October of 1992, with reconstruction scheduled to begin in
early 1993. Five sections of the new design of I-880 will be
constructed of structural steel.
While a mile-long section of the Cypress Street overpass
structure of I-880 collapsed, buildings of various types and
vintage right next to the collapsed freeway experienced very
little or no damage. It is of further interest that the collapsed
portion of I-880 is located on man-made ground, whereas the
surviving elevated section is located on a competent sand
formation.
10
The collapse of a section of the San Francisco-Oakland Bay
Bridge greatly impacted bay area commuting. The Bay
Bridge carries an average of 250,000 vehicles per day be-
tween San Francisco and the cities of the East Bay. The Bay
Bridge is a double-decked steel bridge about 8.5 miles long.
Its west bay crossing is a suspension span, while the east bay
crossing consists of deck trusses and through trusses. About
two miles west of the Oakland toll plaza, 50-foot horizontal
spans, situated along the top and bottom decks and located
above a main pier, serve to link the bridges deck-truss section
to the east with its through-truss section to the west. The
anchor bolts that attached the bridges deck-truss section to
the pier were the only constraint that prevented the two
deck-spans from displacing longitudinally with the bridges
deck-truss section to the east. During the earthquake, large
longitudinal and lateral seismic forces caused these bolts to
fail in shear. Following this failure, the earthquake-induced
longitudinal deformations along the length of the deck-truss
section were sufficiently large (7 in.) to result in collapse of
the upper and lower spans. The cause of the bridge failure was
easily understood, and the 50-foot section was repaired in one
month and opened to traffic again.
11
Landers, California Earthquake
On June 28, 1992 a magnitude 7.5 earthquake, epicentered
near Landers, California in the Southern Mojave Desert,
occurred at 4:58 a.m. At 8:04 a.m. a second earthquake, of 6.5
magnitude and centered near Big Bear Lake 20 miles to the
West of Landers in the San Bernardino mountains, occurred.
Both earthquakes occurred near the so-called Big Bend
of the San Andreas Fault, causing scientists to speculate about
the possibility of a larger earthquake on this conspicuously
quiet stretch of the longest fault in California.
Accelerations of as much as 1.0g were recorded in Lucerne
Valley, and 0.55g in Big Bear, although most epicentral
stations reported peak accelerations of 0.3g or less.
12
Caltrans
had instrumented one of the tall (70 ft.) bridge columns on
Interstate 10, near the city of Colton, after retrofitting the
concrete column with a steel plate jacket as a result of the
1989 earthquake. Although the Landers earthquake showed a
ground acceleration of only 0.1g, acceleration at the top of
the column was 0.8g in the longitudinal direction and 1.02g
in the transverse direction. The column experienced no dam-
age which can be attributed in part to the retrofit method of
wrapping the concrete column in steel.
The Landers earthquake sequence appears to have oc-
curred in a northerly northwest direction, striking the Camp
Rocks, Emerson and Johnson Valley faults. It appears that the
Big Bear seismic event was initiated by movement on the
Camp Rock-Emerson Fault. Two sets of 500kV and two sets
of 220kV transmission lines crossed the Camp Rock-Emer-
son Fault near the north end of the rupture. The fault passed
directly between the legs of a bolted steel frame 220kV tower,
moving two of the legs approximately 9 feet. This movement
resulted in substantial deformation of the steel tower and
failure of several braces. No damage was sustained by the
lines or ceramic insulators and the tower continued to provide
adequate support until repaired.
13
EARTHQUAKE LEGISLATION
Earthquake Hazard Reduction
A review of the history of seismic code development in the
United States
14
helps to more fully understand the lethargy in
bringing modern seismic code requirements into the building
codes. Much like other areas of the world the early seismic
design codes in the United States were the result of disastrous
earthquakes, primarily in California. Major milestones in
seismic code development closely follow many of our signifi-
cant earthquakes. Californias Earthquake Reduction Act of
1986 was signed into law shortly after the 1985 Mexico
Earthquake.
15
This Act is sponsored by the Seismic Safety Commission
which has the responsibility of preparing and administering
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 58
the California Earthquake Hazard Reduction Program. It is a
multidisciplinary commission consisting of 17 commission-
ers and 12 staff. The commissions goal is to significantly
reduce earthquake risk in California by the year 2000. The
responsibility for meeting this goal must be shared by State,
City, and County agencies as well as the private sector.
The first step is an advisory document which is based on
initiatives to improve seismic safety. Between 1987 and 1992
there were 72 initiatives passed by the legislature, with an-
other 42 initiatives scheduled for the 199296 period. The
advisory document contains 150 milestones to measure pro-
gress and record accomplishments.
The hazard reduction program is based on five criteria: (1)
lives saved, (2) damage reduction, (3) socioeconomic conti-
nuity, (4) opportunity (ease of implementation), (5) cost.
These priorities must pass the common sense test of will the
decision maker and the general public consider the initiative
as being practical, sensible, and feasible? The program has 42
initiatives integrating actions needed in the public and private
sector.
The size of the earthquake and the amount of damage
greatly influence safety legislation. For instance, during the
198788 session of the California legislature the Whittier,
California earthquake, M 5.9, occurred with 23 seismic safety
bills being introduced and 11 of the bills passed by the
legislature but only six bills finally being signed into law by
the Governor.
Two years later during the 198990 legislative session the
Loma Prieta earthquake of M 7.1 occurred with 443 bills
being introduced. Of these bills, 164 passed the legislature
and 137 of those were signed into law.
If a historical comparison of legislative action is made for
the 19061989 period, 112 seismic safety bills were signed
into law during that 83-year period. But from 1990 to present,
a short two-year period, 206 seismic safety bills were made
law. Obviously the impact of the Loma Prieta earthquake.
How seismic safety translates into dollars is shown in
Table 3. Out of $670 million for fiscal year 199192 the bulk
of the money, 63 percent, went to the California Department
of Transportation (Caltrans), again as the result of the major
road and bridge damage inflicted by the Loma Prieta earth-
quake.
A review of the 12,500 California State highway bridges
after the 1971 Sylmar California earthquake (6.6 Richter
magnitude) showed that ten percent of the bridges constructed
prior to 1971 would need to be strengthened. The initial
portion of the 1973 program was aimed at retrofitting bridge
hinges. Inexpensive joint restraining devices were developed
and installed. This portion of the program focused on 1,249
bridges statewide and was scheduled to be completed in 1988.
Despite these safeguards, the possibility of bridge damage
was not eliminated. The 1987 earthquake on the Whittier
Fault verified the action Caltrans started in 1973 because
although there was the expected damage during the earth-
quake no bridges collapsed. The program is currently focused
on wrapping a steel reinforcement shield around the concrete
column in bridges with single-column designs.
Seismic Design Provisions
It is impossible to predict the location and magnitude of
earthquakes accurately. It is therefore essential to adopt a
preventive design philosophy in order to avoid repeating
errors in rebuilding after an earthquake and in planning new
construction. This goal is best achieved through code adop-
tion where state-of-the-art seismic design criteria is specified.
One such specification is the AISC Seismic Provisions for
Structural Steel Buildings. First published in 1990 for Load
& Resistance Factor Design, it has been updated in a 1992
version to encompass both LRFD and ASD design proce-
dures.
16
A significant change in the 1992 edition of the seismic
provisions is the conversion to the loads and design format
recommended by the 1991 National Earthquake Hazards
Reduction Program (NEHRP) document.
1
Whereas the provisions contained in the AISC seismic
document are to be used in conjunction with the AISC Load
& Resistance Factor Design (LRFD) Specification, the load
provisions have been modified from those in the LRFD in
order to be consistent with the load provisions contained in
the BOCA, SBCCI Codes, and the ASCE 7-93 Minimum
Design Loads for Buildings and Other Structures.
17
All these
new seismic load provisions are modeled after the 1991
NEHRP earthquake provisions.
Table 2.
California at Risk 19921996
Category
Number of
Initiatives Focus
1
2
3
4
5
20
5
9
5
3
Existing Facilities
New Facilities
Emergency Management
Disaster Recovery
Research and Education
Table 3.
California Seismic Safety Activities
FY 19911992, $670 Million Total
Agency $ Millions
Department of Conservation
Office of Emergency Services
General Services
Public Utilities Commission
Seismic Safety Commission
Department of Water Resources
University of California
Office Statewide Health Planning & Development
Department of Transportation (Caltrans)
6.6
192.6
17.0
0.6
1.4
1.5
3.4
16.6
420.3
SECOND QUARTER / 1993 59
The requirements for analysis and design of buildings
under the 1991 NEHRP and the 1992 AISC Seismic Provi-
sions are based on a seismic hazard criteria, Table 4, that
reflects the relationship between the use of the building and
the level of earthquake to which it may be exposed. This
relationship primarily reflects concern for life safety and,
therefore, the degree of exposure of the public to the hazard
based on a measure of risk.
The purpose of the NEHRP seismic ground acceleration
maps and corresponding seismic hazard exposure groups is
to provide the means for establishing the measure of seismic
risk/performance for a building of any use group, and in any
area of the United States, Table 5.
Seismic performance design requirements get progres-
sively more stringent as the categories proceed from A
through E. The seismic hazard exposure groups listed in
Table 4 are defined in detail, with examples of buildings in
each type, in ASCE 7-93.
17
The most frequently used load combinations given in the
LRFD Specification are repeated in the AISC Seismic Provi-
sions publication in order to reduce the amount of cross-
referencing by the engineer. The load combinations, Table 6,
have been modified to be consistent with the anticipated
ASCE 7-93 document.
The most notable modification is the reduction of the load
factor applied to the earthquake load, E, to 1.0. This results
from the limit states load model used in ASCE 7-93. The
earthquake load and load effects E in the ASCE 7-93 are
composed of two parts. E is the sum of the seismic horizontal
load effects and one half of A times the dead load effects. The
second part adds an effect simulating vertical accelerations
concurrent to the usual horizontal earthquake effects. An
amplification factor to earthquake load E of 0.4R is pre-
scribed. The amount of this amplification was assumed to be
two times the deflections generated by forces specified for a
building with R = 5. This amplification factor is thus 2R / 5
or 0.4R. The added complication that would be required to
consider orthogonal effects with the amplified force is not
deemed necessary.
Base Shear and the R Factor
The equivalent lateral force procedure for a Special Moment
Resisting Frame is greatly influenced by the R or R
w
factor, a
numerical coefficient commonly referred to as a response
modification factor.
NEHRP UBC
V = C
s
W V = ZICW / R
w
C
s
=
1.2A
v
S
RT
2
3
C =
1.25S
T
2
3
For Map Area 7, A
v
= 0.4 Seismic Zone 4, Z = 0.4
Soil/Site Coefficient Importance Factor I = 1.0
S = 1.0 Site Coefficient S = 1.0
V =
1.2(0.4)SW
8T
2
3
V =
0.4(I)1.25(S)W
(12)T
2
3
V = 0.06W V = 0.04W (for equal T values)
Basically, the 50 percent difference in the base shear values
is due to the different response modification factor, R, used
by the Uniform Building Code and NEHRP. The R value
depends on the degree to which the system can be allowed to
go beyond the elastic range, its energy dissipation in so doing,
and the stability of the vertical load carrying system during
inelastic response due to maximum expected ground motion.
It is recognized that the assigned R values must be peri-
odically reviewed as earthquake performance is observed and
more data on material and system performance becomes
available.
18
Under NEHRP design provisions, the design of a structure
Table 4.
Seismic Hazard Exposure Groups
Group III Buildings having essential facilities that are
necessary for post-earthquake recovery and
requiring special requirements for access and
functionality.
Group II Buildings that constitute a substantial public hazard
because of occupancy or use.
Group I All buildings not classified in Groups II and III.
Table 5.
Seismic Performance Categories
Value of A
v
Seismic Hazard Exposure Group
I II III
0.20 A
v
< 0.20
0.15 A
v
< 0.20
0.10 A
v
< 0.15
0.05 A
v
< 0.10
0.20 A
v
< 0.05
D
C
C
B
A
D
D
C
B
A
E
D
C
C
A
Table 6.
Load Combinations
1.4D (3-1)
1.2D + 1.6L + 0.5(Lr or S or R) (3-2)
1.2D + 1.6(Lr or S or R) + (0.5L or 0.8W) (3-3)
1.2D + 1.3W + 0.5L + 0.5(Lr or S or R) (3-4)
1.2D 1.0E + 0.5L + 0.2S (3-5)
0.9D (1.0E or 1.3W) (3-6)
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 60
(sizing of members, connections, etc.) is based on the internal
forces resulting from a linear elastic analysis using the pre-
scribed forces. It assumes that the structure as a whole, under
the prescribed forces, will not deform beyond a point of
significant yield. The elastic deformations then are amplified
to estimate the real deformations in response to the design
ground motion.
1
Earthquake load combinations in the AISC Provision
16
are:
1.2D + 0.5L + 0.2S 0.4R E (3-7)
0.9D 0.4R E (3-8)
The amplification factor (3R
w
/ 8) was derived by using the
similar assumptions that were used in deriving the factor for
ASCE 7-93. The same building type with R = 5 in ASCE 7-93
has a Structural System Coefficient R
w
= 8 in the 1991
Uniform Building Code. The deflection determined by this
R
w
was used as the value to be amplified by 3. Thus (3R
w
/
8)E.
Drift Limits
Model Codes and resource documents such as NEHRP con-
tain specific seismic drift limits, but there are major differ-
ences among them, i.e., UBC drift allowable is
1

3
greater than
that allowed by NEHRP for a Special Moment Frame in steel,
Seismic Hazard Exposure Group I for All other buildings
category, Table 7.
19
There are many reasons for controlling story drift in a
building. Stability considerations dictate that flexibility be
controlled. The stability problem is resolved by limiting the
drift of the building columns and the resulting secondary
moments commonly referred to as P- effects. Buildings
subject to earthquakes also need drift control in order to limit
damage to partitions, emergency stair towers, exterior curtain
walls and other fragile nonstructural elements. The design
story drift limits of NEHRP take into account these needs, and
in order to provide a higher performance standard for essen-
tial facilities the drift limit for Seismic Hazard Exposure
Group III is more stringent than that for Groups I and II, Table
4 and Table 8.
The story drift limitations of ASCE 7-93 and NEHRP
provisions are applied to an amplified story drift that esti-
mates the story drift that would occur during a large earth-
quake. For determining the story drift the deflection deter-
Table 7.
Comparison of 1991 NEHRP and 1991 UBC Drift Limits
Single Story Buildings (Assumed to have a C = 2.75 [UBC] and a C
s
= 2.5 A
a
/ R [NEHRP])
[Z = A
v
]
Framing System
NEHRP UBC
Force
Amplifier
UBC Drift UBC Drift NEHRP Allowable Elastic Drift Ratio of NEHRP to UBC
(0.005h or
[0.04 / R
w
]h)
Scaled to NEHRP
by 0.91R
w
/ R (Delta / C
d
)
SHEG I SHEG II SHEG III
Cd R
R
w
0.91R
w
/ R
I = 1.0 I = 1.25 I = 1.0 I = 1.25
0.025
Hsx
0.020
Hsx
0.015
Hsx
0.010
Hsx NA
0.020
Hsx
0.015
Hsx
Bearing Wall System
Light framed w / sp
CBF
4
3.5
6.5
4
8
6
1.12
1.37
0.0050
0.0050
0.0040
0.0040
0.0056
0.0068
0.0045
0.0055
0.0063
0.0071
0.0050
0.0057
0.0038
0.0043
0.0025
0.0029
1.12
1.05
0.84
0.78
Building Frame System
EBF
Light Framed w / sp
CBF
4
4.5
4.5
8
7
5
10
9
8
1.14
1.17
1.46
0.0040
0.0044
0.0050
0.0032
0.0036
0.0040
0.0046
0.0052
0.0073
0.0036
0.0042
0.0058
0.0063
0.0056
0.0056
0.0050
0.0044
0.0044
0.0038
0.0033
0.0033
0.0025
0.0022
0.0022
1.37
1.07
0.76
1.03
0.80
0.57
Moment Resisting
Frame System
SMF Steel
OMF Steel
SMF Conc.
IMF Conc.
OMF Conc.
5.5
4
5.5
3.5
2
8
4.5
8
4
2
12
6
12
8
5
1.37
1.21
1.37
1.82
2.28
0.0033
0.0050
0.0033
0.0050
0.0050
0.0027
0.0040
0.0027
0.0040
0.0040
0.0046
0.0061
0.0046
0.0091
0.0114
0.0036
0.0049
0.0036
0.0073
0.0091
0.0045
0.0063
0.0045
0.0071
0.0125
0.0036
0.0050
0.0036
0.0057
0.0100
0.0027
0.0038
0.0027
0.0043
0.0075
0.0018
0.0025
0.0018
0.0029
0.0050
1.00
1.03
1.00
0.78
1.10
0.75
0.77
0.75
0.59
0.82
Average 1.03 0.77
Avg. for all moment frames 0.98 0.74
SECOND QUARTER / 1993 61
mined using the earthquake forces E is amplified by a deflec-
tion amplification factor, C
d
(5
1

2
for a SMF of steel) which is
dependent on the type of building system.
The 1991 Uniform Building Code
7
drift provisions are
numerically specific and require that story drift shall be
calculated including the translational and torsional deflection
resulting from the application of unfactored lateral forces.
There are no drift limits on single-story steel-framed struc-
tures with low occupancies.
The AISC Seismic Provisions do not specify specific drift
limits but defer to the governing design code, stating that the
story drift shall be calculated using the appropriate load
effects consistent with the structural system and method of
analysis.
Ordinary Moment Frames
Ordinary moment frames (OMF) of structural steel are moment
frames which do not meet the requirements for special design
and detailing required of the Special Moment Frame. OMF of
structural steel exist in all areas of seismic activity throughout
the country, and experience has shown that this type of building
has responded without significant structural damage.
The 1992 AISC Seismic Provisions for OMF have beam-
to-column joint requirements that allow the use of either fully
restrained (FR) or partially restrained (PR) connections, con-
trary to the Uniform Building Code. But the beam-to-column
connection must meet one of three criteria depending on
whether it is a fully restrained (FR) or partially restrained (PR)
connection:
1. If fully restrained then the connection may conform to
the requirements for SMF except that the required
flexural strength of a column-to-beam joint is not re-
quired to exceed the nominal plastic flexural strength of
the connection
2. If fully restrained with a connection design strength
meeting the requirements of Load Combinations 3-1
through 3-8
3. If either FR or PR connections meeting all the following:
a. The design strengths of the members and connections
shall have a design strength to resist Load Combina-
tions 3-1 through 3-6.
b. The connections have been demonstrated by cyclic
tests to have adequate rotation capacity at a story drift
calculated at a horizontal load of 0.4R E.
c. The additional drift due to PR connections shall be
considered in design.
The provision requiring a demonstration of rotation capac-
ity is included to permit the use of connections not permitted
under the provisions for SMF, such as top and bottom angle
joints, in areas where the additional drift is acceptable.
Column Strength
As the result of the reduction in the actual lateral forces for
use in a code elastic analysis of the structure, overturning
forces are underestimated and are amplified by unaccounted-
for concurrent vertical accelerations. These two load combi-
nations account for these effects:
Axial compression loads:
1.2P
D
+ 0.5P
L
+ 0.2P
S
+ 0.4R P
E

c
P
n
(6-1)
where the term 0.4R is greater or equal to 1.0.
Axial tension loads:
0.9P
D
0.4R P
E

t
P
n
(6-2)
where the term 0.4R is greater or equal to 1.0.
Table 8.
Tentative Allowable Story Drift
Building
Seismic Hazard Exposure Group
I II III
Single-story buildings without equipment
attached to the structural resisting system
and with interior walls, partitions, ceilings,
and exterior wall system that have been
designed to accommodate the story drifts.
No limit 0.020h
sx
0.015h
sx
Buildings with four stories or less with interior
walls, partitions, ceilings, and exterior wall
system that have been designed to
accommodate the story drifts.
0.025h
sx
0.020h
sx
0.015h
sx
All other buildings. 0.020h
sx
0.015h
sx
0.010h
sx
Where hsx is the story height of the story drift calculated.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 62
These load combinations are to be applied without considera-
tion of any concurrent flexure forces on the member.
Column Splices
Column splices, as a minimum, must be able to transmit the
prescribed design code forces, but more stringent provisions
are required for column splices in frames that due to seismic
forces are required to transmit net tension forces.
The AISC Seismic Provisions require partial penetration
welded joints that are subject to net tension to be designed for
forces in excess of the code forces (150 percent of the required
strength) and that the column splice be located three feet from
the beam-to-column connection.
For column splices in seismic design, using either complete
or partial penetration welded joints, beveled transitions as
given in AWS D1.1, Section 9.20,
20
are not required when
changes in thickness and width of flanges/webs occur.
The possibility of developing high tensile stresses in partial
penetration welded column splices during a maximum prob-
able seismic event is real and the use of splice plates welded
to the lower part of the column and bolted to the upper part
should be considered.
The designer should always review the conditions found in
columns in tall stories, large changes in column sizes at the
splice, or where the possibility of a single curvature exists on
a column over multiple stories to determine if special design
strength or special detailing is necessary at the splice.
Panel Zone Design
Cyclic tests of beam-to-column joints has shown the ductility
of shear yielding in column panel zones.
21
The usual Von
Mises shear limit of F
y
/ 3 did not accurately predict the
actual panel zone behavior. Tests have shown that strain
hardening and other phenomena have enabled panel zone
shear strengths in excess of 1.0F
y
dt to be developed.
In calculating the required panel zone shear strength for
AISC LRFD Seismic Provisions, the typical Load Combina-
tions 3-5 and 3-6 are used with the nominal web shear strength
defined as 0.6F
y
dt. In order to provide the same level of safety
as determined by tests and as contained in the 1991 Uniform
Building Code, a lower resistance factor of 0.75 was selected:

v
V
n
= 0.6
v
F
y
d
c
t
p

1 +
3b
cf
t
cf
2
d
b
d
c
t
p

(8-1)
where for this case
v
= 0.75
where:
t
p
= Total thickness of panel zone including doubler
plates, in.
d
c
= Overall column section depth, in.
b
cf
= Width of the column flange, in.
t
cf
= Thickness of the column flange, in.
d
b
= Overall beam depth, in.
F
y
= Specified yield strength of the panel zone steel, ksi.
Eccentrically Braced Frames (EBF)
Whereas concentrically braced frames (CBF) are braced sys-
tems whose worklines essentially intersect at points with no
eccentricities, the EBF is composed of columns, beams, and
braces in which at least one end of each bracing member
connects to a beam at a short distance (eccentricity) from a
beam-to-column connection.
Research
22
has shown that buildings using the EBF system
possess the ability to combine high stiffness in the elastic
range together with excellent ductility and energy dissipation
capacity in the inelastic range. In the elastic range, the lateral
stiffness of an EBF system is comparable to that of a CBF
system, particularly when short link lengths are used.
In the inelastic range, EBF systems provide stable, ductile
behavior under severe cyclic loading, comparable to that of a
SMF system. The design purpose of an EBF system creates a
system that will yield primarily in the links. The special
provisions for EBF systems are intended to satisfy this crite-
rion and to ensure that cyclic yielding of the links can occur
in a stable manner.
Upon publication of the first research report
22
on EBF,
several important applications of this design concept were
employed in the design of major buildings. Ten years later the
Structural Engineers Association of California developed
recommended seismic design provisions for the EBF which
were accepted for inclusion into the 1988 Uniform Building
Code. It is to be noted that the SEAOC and UBC design
provisions for EBF are essentially identical and are based on
the allowable stress design approach, whereas the NEHRP
and AISC Provisions are based on the strength design ap-
proach.
Eccentrically braced frames are designed so that under
earthquake loading, yielding will occur primarily in the links.
The diagonal braces, the columns, and the beam segments
outside of the links are designed to remain essentially elastic
under the maximum forces that can be generated by the fully
yielded and strain hardened links.
EBF have become a well established structural steel system
for seismic resistant construction. Sustained research since
1975 combined with experience from many buildings that
employed the system has provided the database for proper
design of eccentrically braced frames. The most recent EBF
code provisions are contained in the 1992 AISC Seismic
Provisions.
14
This document represents the most up-to-date
and comprehensive code requirements for EBFs currently
available in the United States.
Conclusion
Several years ago, it was not uncommon for local jurisdiction
to each have their own building code. However, this did little
SECOND QUARTER / 1993 63
to promote efficient construction, nor to promote uniform
safety. Today the system has evolved to where most cities
adopt one of three model codes: the Uniform Building Code,
promulgated by the International Conference of Building
Officials (ICBO) and used throughout the western United
States; the National Building Code, promulgated by the
Building Officials and Code Administrators International
(BOCA) and used in the northeastern United States; the
Standard Building Code, promulgated by the Southern Build-
ing Code Congress International (SBCCI) and used in the
southeastern United States.
Since 1957 the Seismology Committee of the Structural
Engineers Association of California has published its seismic
design recommendations. They have also acted as an effective
bridge between seismic research and the application of their
recommendations by assuring that the provisions were
adopted into the UBC in a timely manner. The last major
rewrite of the SEAOC recommendations occurred in 1988,
which formed the basis for the seismic provisions in the 1988
UBC. The SEAOC Seismology Committee is beginning the
preparation of a code change to convert the seismic provisions
in the UBC to a limit state design basis. Their goal is for a
completion time to allow the changes to be incorporated in
the 1997 UBC.
Whenever possible BOCA and SBCCI prefer adopting
design standards by reference.
Unfortunately, the national seismic standard adopted was
ANSI A58.1 / ASCE 7, which was made up of UBC criteria
that was developed by SEAOC. The delay that results from
this technology transfer resulted in the 1987 NBC and 1988
SBC being based on 18 and 14 year old SEAOC recommen-
dations respectively.
But in 1991 BOCA approved seismic code changes based
on NEHRP provisions from its 1988 publication and updated
that to the 1991 NEHRP provisions in 1992. The SBCCI
followed a similar path to code update and the 1993 Standard
Building Code Supplement will contain seismic provisions
based on the 1991 NEHRP.
Within a few months of publication of the June 1992 AISC
Seismic Provisions for Structural Steel Buildings,
16
both
BOCA and SBCCI approved the provisions which will be
referenced in the 1993 NBC, and will appear in the 1993
Amendments to the SBC. Could uniformity in code seismic
design criteria be just around the corner for the United States?
The provisions contained in the AISC Seismic Provisions
for Structural Steel Buildings,
14
are to be used in conjunction
with the AISC LRFD Specification in the design of buildings
in the areas of moderate, high seismicity. The First Edition of
the LRFD Specification was published in 1986. It did not
contain seismic design criteria.
Load and Resistance Factor Design (LRFD) is an improved
approach to the design of structural steel for buildings. The
method involves explicit consideration of limit states, multi-
ple load and resistance factors, and implicit probabilistic
determination of reliability. The designation LRFD reflects
the concept of factoring both loads and resistance. The LRFD
method was devised to offer the designer greater flexibility,
more rationality of design, and possible overall economy.
REFERENCES
1. BSSC, NEHRP (National Earthquake Hazards Reduction
Program), Recommended Provisions for the Development
of Seismic Regulations for Buildings, Building Seismic
Safety Council, Federal Emergency Management
Agency, Washington, DC, 1992.
2. EQE Engineering, Inc., The Performance of Steel Build-
ings in Past Earthquakes, American Iron and Steel Insti-
tute, Washington, DC, 1991.
3. Krawinkler, H., and E. Martinez-Romero, 1989, Per-
formance Evaluation of Steel Structures in Mexico City,
Lessons Learned from the 1985 Mexico Earthquake,
Earthquake Engineering Research Institute.
4. Valle-Calderon, E. D. A. Foutch, and D. K. Hejelmstad,
1989, Investigation of Two Buildings Shaken During the
19 September 1985 Mexico Earthquake, Lessons
Learned from the 1985 Mexico Earthquake, Earthquake
Engineering Research Institute.
5. Etheredge E. and Porcella, R., Strong Motion Data from
the October 1, 1987 Whittier Narrows Earthquake, Open-
file report 87-616, US Geological Survey, October 1987.
6. Pantelic, J. and Reinhorn, A., Report on the Whittier-Nar-
rows, California Earthquake of October 1, 1987, Techni-
cal Report NCEER-87-0026, National Center for Earth-
quake Engineering Research, Buffalo, New York,
November 1987.
7. ICBO, Uniform Building Code, International Conference
of Building Officials, Whittier, CA, 1988 and 1991,
8. Hamburger, R. O., D. L. McCormick, and S. Hom, Sep-
tember-October 1990. Building for Earthquake Survival,
A Historic Perspective, Modern Steel Construction,
AISC, Chicago, IL.
9. U.S. Geological Survey, November 1988, Preliminary
Report of Strong Ground Motion Data, October 17, 1989
Loma Prieta Earthquake, Menlo Park, CA.
10. EQE Engineering, October 1989, The October 17, 1989
Loma Prieta Earthquake: A Quick Look Report, San
Francisco, CA.
11. Dames & Moore, 1989, A Special Report on the October
17, 1989 Loma Prieta Earthquake, Los Angeles, CA.
12. Dames & Moore, Earthquake Engineering News, Volume
No. 4, Summer 1992.
13. EQE International, The Landers and Big Bear Earth-
quakes of June 28, 1992, San Francisco, CA.
14. Martin, H. W. 1993, Recent Changes to Seismic Codes
and Standards: Are They Coordinated or Random
Events?, US National Earthquake Conference, Memphis,
TN, May 1993.
15. Cluff, L. S. 1992, California Earthquake Hazard Reduc-
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 64
tion Program, SEAOC Convention, Ixtapa, Mexico, Sep-
tember, 1992.
16. AISC, Seismic Provisions for Structural Steel Buildings,
June, 1992, American Institute of Steel Construction,
Chicago, IL.
17. ASCE 7-93, Minimum Design Loads for Buildings and
Other Structures, American Society of Civil Engineers,
New York, NY, 1993.
18. SEAOC, Recommended Lateral Force Requirements,
Seismology Committee, Structural Engineers Associa-
tion of California, Los Angeles, CA, 1988.
19. Martin, H. W. 1992, Correspondence on TS-6 Committee
actions on update of 1991 NEHRP Provisions.
20. AWS, D1.1-92, Structural Welding Code, American
Welding Society, Inc., Miami, FL, 1992.
21. Slutter, R., Tests of Panel Zone Behavior in Beam Col-
umn Connections, Lehigh University, Report No.
200.81.403.1, Bethlehem, PA, 1981.
22. Roeder, C. W. and Popov, E. P., Eccentrically Braced
Frames for Earthquakes, Journal of the Structural Divi-
sion, Vol. 104, No. 3, American Society of Civil Engi-
neers, March 1978.
SECOND QUARTER / 1993 65
Although there are seven metric base units in the SI system,
only four are currently used by AISC in structural steel design.
These base units are listed in the following table.
Quantity Unit Symbol
length
mass
time
temperature
meter
kilogram
second
celcius
m
kg
s
C
Similarly, of the numerous decimal prefixes included in the
SI system, only three are used in steel design.
Prefix Symbol
Order of
Magnitude Expression
mega
kilo
milli
M
k
m
10
6
10
3
10
3
1,000,000 (one million)
1,000 (one thousand)
0.001 (one thousandth)
In addition, three derived units are applicable to the present
conversion.
Quantity Name Symbol Expression
force
stress
energy
newton
pascal
joule
N
Pa
J
N = kg m / s
2
Pa = N / m
2
J = N m
Although specified in SI, the pascal is not universally
accepted as the unit of stress. Because section properties are
expressed in millimeters, it is more convenient to express
stress in newtons per square millimeter (1N / mm
2
= 1 MPa).
This is the practice followed in recent international structural
design standards, including the International Standards Or-
ganization (ISO), Draft International Standard for Steel De-
sign,
1
as well as the April 1990 draft of Eurocode 3, Design
of Steel Structures, Part 1General Rules and Rules for
Buildings. It should be noted that the joule, as the unit of
energy, is used to express energy absorption requirements for
impact tests. Moments are expressed in terms of N m.
The following conversion factors relate traditional U.S.
units of measurement to the corresponding SI units:
Multiply by: to obtain:
inch (in.)
foot (ft)
pound-mass (lb)
pound-force (lbf)
ksi
ft-lbf
25.4
305
0.454
4.448
6.895
1.356
millimeters (mm)
millimeters (mm)
kilogram (kg)
newton (N)
N / mm
2
joule (J)
Note that fractions resulting from metric conversion should
be rounded to whole millimeters. Following are common
fractions of inches and their metric equivalent.
Fraction, in. Exact conversion, mm Rounded to: (mm)
1

16
1

8
3

16
1

4
5

16
3

8
7

16
1

2
5

8
3

4
7

8
1
1.5875
3.175
4.7625
6.35
7.9375
9.525
11.1125
12.7
15.875
19.05
22.225
25.4
2
3
5
6
8
10
11
13
16
19
22
25
Bolt diameters are taken directly from the ASTM Specifi-
cation A325M and A490M rather than converting the diame-
ters of bolts dimensioned in inches. The metric bolt designa-
tions are as follows:
Designation Diameter, mm Diameter, in.
M16
M20
M22
M24
M27
M30
M36
16
20
22
24
27
30
36
0.63
0.79
0.87
0.94
1.06
1.18
1.42
The yield strengths of structural steels covered in the metric
LRFD Specification are taken from the metric ASTM Speci-
fications. It should be noted that the yield points are slightly
different from the traditional values.
SI Units for Structural Steel Design
AMERICAN INSTITUTE OF STEEL CONTRUCTION, INC.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 66
ASTM Designation
Yield stress,
N / mm
2
Yield stress,
ksi
A36M 250 36.26
A572M Gr. 345
A588M
345 50.04
A852M 485 70.34
A514M 690 100.07
On the basis of the above selection of units and conversion
factors, the 1986 LRFD Specification has been translated into
the SI system. When necessary, formulas were revised to
make all coefficients nondimensional. In most instances, this
could be achieved by explicitly showing the modulus of
elasticity, E, in the formulation.
The converted LRFD Specification is offered to the federal
agencies and consultants as an interim document to facilitate
design of metric demonstration projects. It will also serve as
an introduction of the SI units of measurement to the general
design profession and fabricating industry. More complete
information is available in the Metric Guide for Federal
Construction, First Edition, prepared by the Construction
Subcommittee of the Metrication Operating Committee. The
guide is available from the National Institute of Building
Sciences in Washington D.C.*
REFERENCES
1. Steel StructuresMaterials and Design, Committee Draft
TC167 / SCI CD10721, ISO, 1991.
* Call (202) 289-7800 for ordering information.
SECOND QUARTER / 1993 67
ABSTRACT
Most designs for buildings with steel frames are based on
girders with simple connections. To eliminate the problems
associated with traditional construction (such as deep and
heavy girders and large deflections during construction)
structural engineers have been searching for a new design
system for a long time. The stub girder system is one example
of such efforts. However, the stub girder system proved to be
uneconomical for most common buildings. A new and differ-
ent approach to composite steel/concrete designs was under-
taken by the writer resulting in light building frames and cost
savings. This new design system is called partial Restraint
Girder System (RGS) (Figure 1). (A composite section is
obtained in buildings with metal deck and concrete floors by
welding steel studs to the top flange.) With RGS two types of
restraint are possible: the first makes use of moment connec-
tions to columns; the second includes concrete reinforcement.
In buildings utilizing composite girders, deflections were
controlled by either shoring, camber, or further increase in
girder size. RGS has arisen as a viable and cost effective
alternative.
In traditional designs, the engineer determined the build-
ings moment diagram from a moment distribution or stress
analysis. In the RGS method, the Structural Engineer can
control the maximum and minimum values of moment on the
moment diagram (the governing design values) from the
outset to fit his design, by establishing the amount of restraint.
Although composite girders with partial restraints improve
the moment resistance of composite girders significantly,
such design is commonly ignored and the codes of practice
give no guidance as to procedures that might take advantage
of the improved properties.
PAST RESEARCH
Extensive knowledge is available on non-composite girder-
to-column moment connections. Reference 3 provides a good
summary with design examples of various non-composite
girder-to-column moment connections. Composite girders
with girder-to-column moment connections is a new topic and
less covered in past research.
Karl Van Dalen (Reference 4), Ammerman, and Leon
(Reference 6) and others, realized the significance of com-
posite girders with negative concrete reinforcement (com-
posite connections). They tested specimens to determine
strength, stiffness, and ductility. Reference 6 provides a good
summary of past research on composite girders with negative
reinforcement (semi-rigid connections).
Also, Ammerman recognized the significance of semi-
rigid composite connections (Reference 9) and suggested a
method for the design of frames incorporating such connec-
tions. By considering the construction phase, RGS improves
on the previous work. This paper describes RGS, provides the
mathematical formulae which describe the system and pro-
vides examples of buildings designed and built with RGS.
INTRODUCTION
The traditional design for buildings with steel frames is based
on composite girders with simple connections. The disadvan-
tage of this traditional design is that the entire moment re-
quirement is at one portion of the girder resulting in large size
girders. Also, girders have large mid-span deflections during
construction when the concrete is wet. In order to eliminate
these disadvantages, the designer specified camber or tempo-
rary shoring. However, since both of these methods are costly
and difficult to implement, contractors often preferred to do
without them and instead, increased girder sizes even further.
With partial restraints, girder sizes can be decreased and the
deflections reduced.
Two different restraint types are possible, as follows:
a. Girder-to-column moment connection.
b. Negative concrete reinforcement.
Neil Wexler is president, N. Wexler, P.E., P.C. Consulting
Structural Engineers.
Composite Girders with Partial Restraints:
A New Approach
NEIL WEXLER
Fig. 1. Composite girder with partial restraint.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 68
The two different types of restraint are best utilized based
on the following rule of thumb: when deflections during
construction are large, and/or the girder sizes are governed by
construction loads, girder-to-column connections are pre-
ferred; when deflections do not govern, and the girder size is
governed by superimposed loads, negative concrete rein-
forcement bars are preferred; however excellent results are
achieved when a combination of both restraint types is used.
Just like a composite girder with simple connections, the
design of a Restraint Girder System is also done in two
phasesconstruction phase, when the concrete is wet and the
final phase, after the concrete hardens.
CONSTRUCTION PHASE
At this phase, the steel girder alone supports all the loads.
Some steel girders with simple connections have significant
mid-span deflections at this phase. Introducing end moment
connections results in reduced mid-span deflections. For ex-
ample, for a beam with fixed connections, Figure 2 shows that
mid-span deflections can be reduced by as much as 58 percent
if only one end is fixed and by 80 percent if both ends are
fixed. These are very significant reductions. Considering that
the end moment connections also have the added benefit of
reducing the mid-span moment, it becomes quite clear what
a powerful design tool has been created.
To provide the rigidity required for this phase, the end
moment connection must be designed as a rigid connection
(AISC Type 1 construction) for this stage (Figure 3). It
provides just enough strength and rigidity to hold the original
angles between the members unchanged. During this phase
all connection components are stressed elastically. The con-
nection is strong enough to hold the original angles between
members unchanged, reducing the center moment and mid-
span deflections.
FINAL PHASE
At this phase, the steel girder acts compositely with the
concrete. The girder now is both strong and rigid.
Once the concrete hardens, superimposed loads such as
partitions, mechanical, ceiling, and live loads are applied. At
this time, the moment at the girder end wants to increase,
however since the connection has reached its elastic capacity,
it will deform plastically. This now corresponds to AISC Type
3 construction; the connection now becomes semi-rigid. All
excess moment shaken-off by the semi-rigid moment con-
nection is now transferred to the middle section of the girder.
Since this middle section is composite with the concrete, it is
both strong and rigid. Therefore, any deflections associated
with the final phase are small.
The design described above results in smaller girders and
reduced costs. Where before A36 steel was used, governed by
deflection considerations, A572 steel now often becomes
more economical.
The sensitivity of the RGS system to deflections when the
concrete is wet and unevenly placed ought to be investigated
by the design engineer for each individual project. In some
cases the RGS system should be specified with recommenda-
Figure 2 Figure 3
SECOND QUARTER / 1993 69
tions for the concrete pour sequence and the acceptable
locations of construction joints.
END-MOMENT CONNECTIONS
A cost efficient end-moment connection is an end-plate con-
nection (for moderate size moments) (Figure 3a and 3b). It
performs well as a rigid connection during the construction
phase, and a semi-rigid connection at the final phase. An end
plate is a particularly good choice because not only does it
deliver forces to the column but it also reinforces the column
by spreading compression forces over larger areas, just like a
bearing plate, thus reducing the need for compression column
stiffeners. It is especially economical when full penetration
welds are not required. However, other connections can also
be used. Figure 3d shows a girder to column moment connec-
tion with angles. Angles are also a good choice because the
bottom flange is reinforced against local buckling by the
horizontal leg of the angle. Reference 3 and others provide
design guidelines for the design of such connections.
Increasing the connection size beyond that which provides
full fixity during the construction phase is usually not neces-
sary and proves to be uneconomical. Therefore, for best
economy, the end moment connection need not be over-de-
signed. Details A and B in Figure 3 show a relatively inex-
pensive moment connection. The connection shown in detail
C is more expensive.
One way to evaluate a girder with moment connections is
by making use of the connection moment rotation curve
(Figure 4). A composite girder with partial restraint behaves
just like a steel girder with full restraint when the concrete is
wet (line A). After the concrete hardens, and additional loads
are superimposed, the connection provides additional re-
straint until yielding; then the girder behaves just like a simple
supported composite girder. Figure 4 shows a moment rota-
tion curve with concrete reinforcement added at the joint. The
connection curves shown are diagrammatic and in reality
yielding may occur sequentially.
In order to determine the various points on the moment
rotation curve, Figure 5 is reproduced herein from Refer-
ence 3.
ADDITIONAL RESTRAINT
Research done by Professors Karl Van Dalen and Hernan
Godoy at Queens University, Kingston, Ontario (Refer-
ence 4) revealed that additional moment strength can be
achieved at the beam-column connection if only 0.46 percent
of the concrete slab area is provided as slab reinforcement.
This additional strength is at least equal to the ultimate
moment capacity of the composite beam and is not influenced
by the type of connection between the steel elements. The
rotational capacity of the composite beamcolumn connec-
tion is also at least equal to that of a conventional, non-com-
posite rigid steel connection.
The AISC specifications for Structural Steel for Buildings
(Reference 2) allows the calculations of the negative design
moment strength based on the plastic stress distribution of the
composite section, provided that the following are met:
a. Shear connectors are located in the negative moment
region.
Fig. 4. Girder-to-column connection + reinforcement.
Fig. 5. Moments and end rotation for various
load/beam conditions.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 70
b. The slab reinforcement is adequately developed.
c. Steel beam is compact and braced.
The designer can use this additional strength to reduce the
girder size further. Only additional studs and negative con-
crete reinforcement are needed. However, in order to ensure
a uniform cracking pattern in the slab in the vicinity of the
column, Karl Van Dalen (Reference 4) recommends that at
least twice the minimum area of steel reinforcement be ex-
tended on each side of the column centerline (Figure 6).
UNBALANCED LOADS
For a long time engineers assumed that unbalanced loads
might overstress non-composite steel girders designed with
partial restraint and therefore avoided the use of such restraint
in steel buildings. RGS however is not very sensitive to
unbalanced loads for two reasons:
a. Traditionally, structures have been analyzed on the basis
of frame actionmeaning beams with joints which
are allowed to rotate when subjected to unbalanced
moments. For building structures with concrete floors an
additional horizontal restraint exists. This restraint is
provided by the concrete diaphragm and is usually ig-
nored by engineers. Unbalanced loads create a joint
rotation; any joint rotation is associated with horizontal
and vertical translations. However, in certain buildings
with girders connected to concrete floor diaphragms
horizontal translations are restrained. Therefore unbal-
anced loads in buildings with concrete floors are gener-
ally less able to generate joint rotation.
b. If the load is increased the reinforcement might yield.
Tests indicate that reinforced concrete has some ductility
to transfer of moments from one section to another after
first yielding of reinforcement. When tensile reinforce-
ment reaches yield at one section it will continue to yield
while the section rotates. For a continuous member, the
load will increase until all sections reach yield or until
the concrete reached ultimate strain. As a result for most
concrete structures moment redistribution is possible
and accepted by the Building Codes.
Because of the above reasons, for most common buildings
using RGS, adequate strength is assured with unbalanced
loads present if adequate concrete reinforcement is provided.
DUCTILITY
Ductility is associated with the ability of the joint to rotate
after yielding. Joint rotation can be prevented by premature
local or overall buckling of the bottom flange and buckling
of the web. The use of under-reinforced sections and class 1
steel shapes assures adequate post-yielding rotations.
COMPOSITE STUDS
Stud-design criteria is similar to composite girders without
restraint, with the exception that if top reinforcement is used
for restraint, then additional studs are required between the
point of maximum negative moment and point of zero mo-
ment. The number of such studs shall be selected to develop
the negative moment (Figure 7).
ALTERNATIVES
Alternatives to composite girders with partial restraint in-
clude the stub girder system, haunch girders, and simple
composite girders. These alternatives however, are expen-
Fig. 6. RGSConcrete reinforcement.
Fig. 7. Composite beam-stud requirements.
SECOND QUARTER / 1993 71
sive, complicated, deep, and might require duct openings,
camber, shoring.
ADVANTAGES
The advantages of composite girders with partial restraint are
many. First, the design usually results in shallow and small
girders with no duct penetrations. The lighter weight girders
also have small mid-span deflections. Camber or shoring are
not required. Wind loads can be incorporated as part of the
33 percent increase in allowable stresses at no additional
costs. Medium to large spans can be accommodated. Long
term creep deflections associated with shored construction
are eliminated. The engineering analysis and design is simple
and suitable for hand calculations or computer use. All engi-
neering principles involved are based on the AISC specifica-
tions and in accordance with standard practice.
It is important to point out what is new about RGS. Girder-
to-column moment connections, rigid or semi-rigid, are not
new; negative concrete reinforcement with composite con-
struction is not new for wind loads, but it is new for gravity
loads when used to reduce the girder size. (Some reports
indicate that negative reinforcement with composite con-
struction has been used with the stub girder system). The use
of rigid and semi-rigid construction in the same frame at the
same location is new. The use of such construction in con-
junction with negative concrete reinforcement is also new.
The traditional use of moment connections alone or the
traditional use of negative reinforcement alone cannot pro-
vide the benefits which are created by the RGS system. For
this reason, RGS has arisen as a viable new alternative system
for building construction. The final result, which is a small
girder system with controlled deflections, and without shor-
ing or camber, is new.
EXAMPLES
1. Capitol Square Office BuildingColumbus, Ohio
(Figure 9)
This is a 28-story office tower with a triangular floor plan.
The bay size is 3030. The floor construction consists of 2-in.
deep metal deck and 3.25-in. light weight concrete. The filler
beams are W1422 at 10 feet on center. The girders are
composite, restrained, hunched girders. Both hunches and
restraints were used in order to develop the maximum possi-
ble negative moment at the columns and reduce the mid-span
moment. The straight portion of the girders was W1430.
Hunches were fabricated from 16-in.-deep sections, cut
diagonally. Restraint was obtained with end plates. This con-
tinuous girder system was also used to provide additional
lateral load resistance.
Contractor: Turner Construction Company, Columbus, Ohio.
Steel Fabricator: Ohio Steel Fabricators
2. Minolta Office BuildingRamsey N.J.
(Figure 8)
A 4-story office building with a steel frame on a concrete
foundation. The building has a horseshoe footprint and
120,000 sq. ft of space. The bay size is 2530 with girders
framing the long direction. 3.25 inches of light-weight con-
crete were poured over 2-in. composite metal deck. 14-in.
deep composite filler beams are spaced at 10 feet on center
and span the 25-ft dimension. 14-in. composite girders with
partial restraint at columns support all gravity and wind loads.
Simply supported girders, 50 ksi steel, would have been
W1453 with a 1-in. camber and 57 studs. Using partial
restraint, only W1443 girders were used with no camber and
24 studs. Many mechanical units and a continuous roof screen
created heavy congestion on an open web joist-framed roof.
Despite the roof congestion, the total steel weight was under
7.5 lbs. per sq. ft. No floor deflections or vibrations were
reported. The contractor and steel fabricator reported easy
fabrication and construction details.
Steel fabricator: Mulach Steel, Pennsylvania
Fig. 8. Part planMinolta Office Building.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 72
3. Parking/Retail BuildingNew York, N.Y.
(Figure 10)
This is a 45,000 sq. ft building with retail on the ground
floor, parking in the basement, and a playground/community
area on the roof. The bay size varies; filler beams are 18 to 22
feet long and the girders 17 to 35.5 feet long.
A traditional design with simply supported composite gird-
ers would have resulted in W1876 or W2462 cambered
girders. Using Partially Restraint Girders resulted in W1645
(F
y
= 50) for the first exterior girder and W1636 (F
y
= 50) for
interior girders without camber. The end-moment connection
is an end plate. The shallower girders were necessary for a
lower overall building height. Figure 11 shows a comparison
Fig. 9. Part planCapitol Square Office Building.
Fig. 10. Part planParking/Retail Building.
Fig. 11. Summary.
Figure 12
SECOND QUARTER / 1993 73
of different design schemes using RGS alternatives. Scheme
C which includes a girder-to-column moment connection and
negative concrete reinforcement is the best design alternative.
It requires no camber, yet the result is small girder sizes,
economical connections, and light concrete reinforcement.
ANALYTICAL EVALUATIONS
Figure 12 shows formulas which govern the design of RGS
using LRFD. Equations 6-1 and 6-2 are based on Van Dalen
(Reference 4). While Van Dalen does not address the case of
a girder-to-column moment connection alone, an increased
amount of minimum reinforcement is recommended by the
writer to control cracking and for unbalanced loads.
Equation 6-3 is based upon partial-partial restraint when
the amount of concrete negative reinforcement required,
based on loads, is less than the maximum that can be provided.
Equation 6-4 represents the maximum amount of concrete
negative reinforcement that can be provided and still ensure
its yielding.
A ductility factor K is provided to ensure that the concrete
reinforcement will yield first. This factor can vary depending
upon certain factors such as the steel girder size, the type of
girder-to-column moment connection and/or the type of bot-
tom flange restraint.
The number of composite studs should be selected based
on the formulas shown in Figure 7.
Figure 13 shows tables prepared by the author for a quick
design of RGS in a engineering office.
ADDITIONAL RESEARCH
It is recommended that resources be allocated for research
into the following and other topics for better prediction of the
behavior of the RGS system in building structures:
1. Requirements for column stiffeners.
2. Behavior under reversed loading. Wind and seismic
loading.
3. Short and long term effects of shrinkage, creep, relaxa-
tion.
4. Non-linear behavior of steel connections and concrete
reinforcement.
5. Pattern loading conditions.
6. Column unbraced length.
Figure 13a Figure 13b
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 74
7. Shored versus unshored construction.
8. Local buckling.
CONCLUSIONS
A new girder system was presented. The system resolves the
problems associated with traditional steel construction. Re-
straint Girder System (RGS) provides a powerful tool for
reducing girder sizes and mid-span deflections. End restraint
is achieved with end-moment connections and/or concrete
top reinforcement. The design is especially economical when
the end-moment connection is detailed with economy in
mind. The shape of the moment diagram can be controlled by
the designer, resulting in cost savings. The new design
method can be used with significant savings for most steel
buildings with metal deck concrete and composite action. The
resulting analysis and design method will most likely become
a favorite for steel design into the next century, creating a new
class of building structures. In the future engineers might find
it practical to apply RGS knowledge to other building struc-
tural components such as filler beams and also to bridges.
ACKNOWLEDGMENTS
The author would like to thank Professor Karl Van Dalen,
Queens University, Kingston, Ontario for his contribution on
the subject of composite connections.
REFERENCES
1. Manual of Steel Construction, Allowable Stress Design,
Ninth Edition, AISC, 1989.
2. Manual of Steel Construction, Load and Resistance Factor
Design, First Edition, AISC, 1986.
3. Blodgett, Omer W., Design of Welded Structures, The
James F. Lincoln Ave Welding Foundation, Cleveland
Ohio, 1966.
4. Van Dalen, Karl and Hernan Godoy, Strength and Rota-
tional Behavior of Composite BeamsColumn Connec-
tions, Queens University, Kingston, Ontario, October 5,
1982.
5. McCormac, Jack C., Structural Steel Design, LRFD
Method, Harper & Row, New York, 1989.
6. Ammerman, Douglas J. and Roberto T. Leon, Behavior of
Semi-Rigid Composite Connections, University of Minne-
sota, AISC Engineering Journal, Volume 24, No. 2, 1987.
7. Fling, Russel S., Practical Design of Reinforced Concrete
Structures, John Wiley & Sons.
8. Building Code Requirements for Reinforced Concrete, ACI
318-89.
9. Ammerman, Douglas James, Behavior and Design of
Frames with Semi-Rigid Composite Connections, Thesis
Degree of Doctor of Philosophy, University of Minnesota.
Figure 13c
SECOND QUARTER / 1993 75
Eng ine e ring
J ourna l
AMERICAN INSTITUTE OF STEEL CONSTRUCTION, INC.
Pa g e 77: W. Sa mue l Ea ste rlin g a n d Lisa
Gon za le z Giroux
She a r La g Effe c ts in Ste e l Te nsion Me mb e rs
Pa g e 90: N. Kishi, W. F. Che n , Y. Goto, a n d
K. G. Ma tsuoka
De sig n Aid of Se mi-rig id Conne c tions for Fra me
Ana lysis
Pa g e 108: Atorod Azizin a min i a n d Ba n g a lore
Pra ka sh
A Te nta tive De sig n Guid e line for a Ne w Ste e l Be a m
Conne c tion De ta il to Comp osite Tub e Columns
3rd Qua rte r 1993/ Volume 30, No. 3
INTRODUCTION
The non-uniform stress distribution that occurs in a tension
member adjacent to a connection, in which all elements of the
cross section are not directly connected, is commonly referred
to as the shear lag effect. This effect reduces the design
strength of the member because the entire cross section is not
fully effective at the critical section location. Shear lag effects
in bolted tension members have been accounted for in the
American Institute of Steel Construction (AISC) allowable
stress design specification
1
(ASD) since 1978. The 1986 load
and resistance factor design specification
2
(LRFD) and the
1989 ASD specification
3
stipulate that the shear lag effects
are applicable to welded, as well as bolted, tension members.
Past research on the subject of shear lag has focused
primarily on bolted tension members. Recently, more atten-
tion has been given to welded members, evident by their
inclusion in the AISC specifications. Shear lag provisions for
welded members were introduced into the specifications pri-
marily because of a large welded hanger plate failure.
8
To
maintain a uniform approach to both welded and bolted
members, the same provisions for shear lag in bolted mem-
bers were applied to welded members. Additional require-
ments for welded plates were added. However, the applica-
tion of the shear lag requirements to welded members has
raised several questions.
This paper examines shear lag in steel tension members in
the following context. First, the background for the current
AISC specification provisions is reviewed. Second, the re-
sults of an experimental research program in which 27 welded
tension members were loaded to failure is presented. Third,
based on the first two parts of the paper, recommended
changes to the AISC specifications are presented.
BACKGROUND FOR CURRENT
DESIGN PROVISIONS
Bolted Connections
The shear lag provisions in the current AISC specifications
2,3
are based on work reported by Chesson and Munse.
6,11
This
work included experimental tests of riveted and bolted ten-
sion members conducted by Chesson and Munse and a review
of experimental tests by other researchers. Chesson and
Munse
6
defined test efficiency as the ratio, in percentage, of
the ultimate test load to the product of the material tensile
stress and the gross area of the specimen, and used this ratio
to evaluate the test results. Several factors influence the test
efficiency of connections failing through a net section: the net
section area, a geometrical efficiency factor, a bearing factor,
a shear lag factor, and a ductility factor.
The data base Chesson and Munse gathered included tests
that failed in a variety of ways, including rupture of the net
section, rivet or bolt shear, and gusset plate shear or tear-out.
However, only tests exhibiting a net section rupture, approxi-
mately 200, were included in the validation of the tension
member reduction coefficients. Munse and Chesson seldom
observed efficiencies greater than 90 percent and therefore
recommended, for design use, an upper limit efficiency of 85
percent.
11
Chesson
5
reported on two additional studies that
recommended maximum efficiencies of 0.75 and 0.85.
Fourteen of the 30 tests conducted by Chesson and Munse
6
failed by net section rupture. Nine of the 14 tests failed at load
levels exceeding the gross cross section yield load. Tests
reported by Davis and Boomslitter
7
were used in the overall
data base and also exhibited net section failures at load levels
exceeding gross section yield. References to other tests are
given by Chesson and Munse.
Research reported prior to 1963 indicated that shear lag
was a function of the connection length
5
and the eccentricity
of the connected parts.
7
Combining previous research results
with their own investigation of structural joints, Munse and
Chesson
11
developed empirical expressions to account for
various factors influencing the section efficiency. The two
most dominant parts of their formulation were the net section
calculation, which accounts for stagger of the fasteners, and
the shear lag effect. The shear lag expression is given by
U = 1
x
_
l
(1)
where
U = shear lag coefficient
x
_
= connection eccentricity
l = connection length
An AISC Task Committee concluded from a review of
Munse and Chessons results that the recommended design
W. Samuel Easterling is associate professor in the Charles E.
Via, Jr. Department of Civil Engineering, Virginia Polytechnic
Institute and State University, Blacksburg, VA.
Lisa Gonzalez Giroux is staff engineer, Hazen and Sawyer,
P.C., Raleigh, NC.
Shear Lag Effects in Steel Tension Members
W. SAMUEL EASTERLING and LISA GONZALEZ GIROUX
THIRD QUARTER / 1993 77
procedure could be simplified.
10
The simplification is in the
form of coefficients given in the AISC Specifications.
2,3
Al-
though the work of Chesson and Munse included the effects
of several factors on the net section efficiency, the AISC
specifications only account for the two dominant factors, net
area and shear lag. The commentaries of both specifications
include Equation 1 as an alternate approach for determining
the shear lag coefficients. The calculation of the effective net
area, A
e
, incorporates the shear lag coefficient and is given by
A
e
= UA
n
(2)
where
A
n
= net area
Welded Connections
In 1931 the American Bureau of Welding published the results
of an extensive study in which safe working stresses for welds
were determined. The American Bureau of Welding was an
advisory board for welding research and standardization of
the American Welding Society (AWS) and the National Re-
search Council Division of Engineering.
4
The study was a
collaborative effort between three steel mills, 39 fabricators,
61 welders, 18 inspectors, and 24 testing laboratories. Several
specimen configurations were used in the test program and
were assigned a series designation, e.g. 2400, 2500, etc.,
based on the configuration. Those directly applicable to this
discussion consist of flat plate specimens, welded either
longitudinally or both longitudinally and transversely. Both
single and double plate tension specimens, as shown in Fig-
ure 1, were tested in the research program.
Most of the tests in the AWS program failed through the
throat of the weld; but several of the specimens ruptured
through the plate. The tests that ruptured are the ones appli-
cable to the study described here. Key results from these tests
have been taken from the report and are presented in Table 1.
Figure 2 is a plot of the results in terms of plate thickness vs.
experimental shear lag coefficient (efficiency), U
e
.
Several trends are apparent in Figure 2. First, as the plate
thickness increases, the scatter in the data tends to increase,
with the average experimental shear lag coefficient increasing
slightly. This trend appears to hold except for the
5

8
-in. group,
which shows the least scatter, although this is the group with
the smallest number of tests.
Second, the amount of scatter in the
3

4
-in. group is unexpect-
edly high. There are groups of tests in which specimens have
virtually identical details, yet the results vary by as much as 30
percent. For instance, consider the two
3

4
-in. specimens in series
2200. The specimen details are nearly the same, yet the experi-
mental efficiency varies from 0.69 to 1.03. Likewise, the
3

4
-in.
specimens of series 2400 had very similar details, but the experi-
mental efficiencies varied from 0.65 to 0.94.
A number of factors may have caused the scatter, including
variation in the quality of the welds. An interesting observa-
tion pertaining to the issue of weld quality was made while
reviewing the AWS report. The last column of Table 1 is a
code used in the report to indicate the welding process (arc or
gas), fabricating shop, welder, and mill that supplied the steel.
Nine specimens, all of which were arc-welded, failed at an
efficiency less that 0.80. Most of these specimens had com-
panion specimens, which had similar fabrication details, yet
they exhibited test efficiencies well above 0.80. A hypothesis
that welding techniques, which may have created gouges or
notches in the base material, caused the scatter in the data was
formed by the authors of this paper. This seems plausible
because the nine tests with efficiencies below 0.80 were
fabricated in two shops, by three welders, using steel from
two mills (3 heats), and seven of those were welded in the
same shop by two welders. Unfortunately, this hypothesis
cannot be confirmed for tests conducted more than 60 years
ago.
The results of the AWS research were considered in the
development of the AISC specification provisions accounting
for shear lag in welded members. However, as will be pre-
Fig. 1. AWS test specimen configuration.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 78
sented later in this paper, questions have arisen regarding the
application of the provisions to welded members. A research
program was initiated to address the questions. The remainder
of this paper presents the results of the research program.
RESEARCH PROGRAM FOR WELDED
TENSION MEMBERS
This section of the paper summarizes a research project
conducted at Virginia Tech focusing on the application of
shear lag specification provisions to welded tension mem-
bers, presenting both experimental and analytical results. The
experimental program included tests of 27 welded tension
members, along with the associated tensile coupon tests.
Analytical studies included elastic finite element analyses of
the experimental specimens, as well as a review of the AISC
specification provisions pertaining to shear lag.
Description of Experimental Specimens
Each test specimen consisted of two members welded back-
to-back to gusset plates, as shown in Figure 3. The gusset
plates were then gripped in a universal testing machine and
pulled until failure. Use of double members minimized the
distortion due to the out-of-plane eccentricity, however,
eccentric effects were ignored in the design of the test
specimens.
Three types of member were tested: plates, angles, and
channels. Fillet weld configurations used for each member
type, except the plates, were longitudinal, transverse, and a
Table 1.
AWS Test Results
a
AWS
Series
t
p
(in.)
w
(in.)
l
(in.)
F
y
(ksi)
F
u
(ksi)
A
g
F
y
(kips)
A
g
F
u
(kips)
Test Load
(kips)
U
e
Proc-Fab-
Weld-Mill
b
2200 0.75 7.5 12.0 36.3 57 204 321 221 0.69 A-Q-A-C
2200 0.75 7.5 12.0 33.2 56.9 187 320 329 1.03 G-AZ-B-I
2400 0.375 7.5 6.0 35.7 58 201 326 248 0.76 A-Q-B-C
2400 0.5 7.5 8.0 37.2 60.2 279 452 406 0.9 A-P-A-C
2400 0.5 7.5 8.0 37 59.2 278 445 303 0.68 A-Q-B-C
2400 0.5 7.5 8.0 39.2 62.2 294 467 382 0.82 G-AZ-B-I
2400 0.75 7.5 12.0 36.5 59.2 411 667 432 0.65 A-C-A-B
2400 0.75 7.5 12.0 36.4 59.6 410 671 484 0.72 A-Q-B-C
2400 0.75 7.5 12.0 33.5 57 377 641 600 0.94 G-AZ-B-I
2500 0.75 4.0 4.0 35.6 60.4 106.8 181.2 170.6 0.94 G-AZ-A-I
2600 0.5 7.5 4.0 37 59.3 139 222 149 0.67 A-Q-A-C
2600 0.75 7.5 8.0 36.5 59.2 205 333 186 0.56 A-C-A-B
2600 0.75 7.5 8.0 36.4 59.6 205 335 200 0.6 A-Q-B-C
2700 0.5 4.0 2.0 36.8 62.1 147.2 248.4 237.6 0.96 G-AZ-A-I
2700 0.75 4.0 4.0 35.6 60.4 213.6 362.4 350 0.97 G-AZ-B-I
2700 0.75 4.0 4.0 35.6 60.4 213.6 362.4 345 0.95 G-AZ-B-I
2800 0.375 7.5 2.0 35.7 58 201 326 282 0.87 A-N-A-C
2800 0.375 7.5 2.0 35.7 58 201 326 239 0.73 A-Q-A-C
2800 0.375 7.5 2.0 37.5 58.2 211 327 278 0.85 G-AZ-B-I
2800 0.375 7.5 2.0 37.5 58.2 211 327 275 0.84 A-CZ-A-I
2800 0.5 7.5 4.0 39.2 62.2 294 467 417 0.89 G-AZ-A-I
2800 0.625 7.5 6.0 36.6 61.6 343 578 500 0.87 A-C-A-B
2800 0.625 7.5 6.0 37.3 57 350 534 475 0.89 A-Q-A-C
2800 0.625 7.5 6.0 33.4 57 313 534 499 0.93 G-AZ-B-I
2800 0.625 7.5 6.0 33.4 57 313 534 520 0.97 A-CZ-A-I
2800 0.75 7.5 8.0 36.4 59.6 411 671 606 0.90 A-N-A-C
2800 0.75 7.5 8.0 36.4 59.6 411 671 590 0.88 A-P-A-C
a. All welds nominally
3
8-in.; measured variation between
3
8 and
1
2-in.
b. Procwelding process; A = arc welding G = gas welding
Fabfabricator designation
Weldwelder designation (within particular fabricating shop)
Millmill designation for steel supply
THIRD QUARTER / 1993 79
combination of both longitudinal and transverse. For the
plates, two different lengths of longitudinal weld and a com-
bination of longitudinal and transverse welds were used.
For a given specimen configuration, three nominally iden-
tical tests were conducted; specimens with only transverse
welds were the exception. Calculations indicate that tension
members connected with only transverse fillet welds will
always fail through the welds. For the purpose of confirming
the calculations three specimens were fabricated with only
transverse welds. Details of the specimens are given in
Table 2. Test designations in Table 2 indicate the type of
member (P = plate, L = angle, C = channel), weld configura-
tion (L = longitudinal, T = transverse, B = longitudinal and
transverse) and specimen number for a given member type
and weld configuration. For instance, test designation P-B-2
is a plate specimen with both longitudinal and transverse
welds and is the second test in that particular group. An
additional number appears in the weld designation for some
of the plate specimens (e.g. P-L2-3). This is because the
longitudinal weld lengths were varied in some of the plate
specimens that were fabricated with only longitudinal welds.
In an attempt to ensure net section failures in the members,
all welds, except the transverse welds, were designed to have
1015 percent greater strength than the gross section tensile
strength of the member. The width and thickness of the
connected member elements prevented oversizing of the
transverse welds. Welds were balanced by size for all angle
specimens, except L-B-1a, with the longitudinal weld lengths
being equal on each specimen. Specimen L-B-1a was unbal-
anced with the two longitudinal welds being the same size
and length.
Strain gages were used in one of the tests for each member
type to study the stress distribution near the critical section of
the member and the distribution of stress in the member along
the length of the connected region. A displacement transducer
was used to monitor the overall cross head movement. This
measurement is only of qualitative value since it includes any
slip between the specimen and the testing machine grips.
Each specimen was whitewashed before testing to permit the
observance of qualitative yield pattern formation. Complete
specimen details are reported by Gonzalez and Easterling.
9
Two aspects of the authors research program should be
kept in mind while reviewing the following results. The first
is that the number of tests was limited, compared to the many
tests available for consideration when the shear lag provisions
were developed by Munse and Chesson. Second, the member
sizes used to fabricate the test specimens were small. The
capacity of the testing equipment available at the time the
tests were conducted limited the member sizes. There are
undoubtedly size effects that the results of this study do not
reflect. However, the same can be said of the data base that
forms the basis for the current shear lag specification provi-
sions, and thus the results of this study can be considered
similar to the bolted and riveted test results.
Description of Analytical Models
Linear elastic finite element analyses were performed for
experimental test specimens using ANSYS, a commercial
finite element analysis package.
12
None of the transverse
welded members were analyzed.
A two-dimensional, four node, isoparametric plane stress
element was used to model the plate specimens. The angles
and channels were modeled using three-dimensional, four
node, quadrilateral shell elements. Linear elastic spring ele-
ments simulated the welds. The spring force constant for the
weld elements was determined using a calibration procedure.
Only the members and the welds were modeled elastically; Fig. 3. Test specimen configuration.
Fig. 2. Plate thickness vs. experimental shear lag
coefficient for AWS tests.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 80
the gusset plates were considered to be rigid. Interface, or gap,
elements were used to prevent the member from displacing
into the gusset plate. A typical finite element mesh and
boundary conditions for a model of an angle specimen are
illustrated in Figure 4. Plate and channel models were con-
structed in a similar manner. Only results from the plate
models are presented in this paper. Results from other analy-
ses are reported by Gonzalez and Easterling.
9
The stiffness for the weld elements was determined by
calibrating a model of a plate with 5-in. longitudinal welds to
the corresponding experimental specimen. In the calibration
process, the model was analyzed with eight different weld
stiffness values, ranging from 100 to 5,000 k/in. The com-
pletely rigid case was also considered. The remaining analyti-
cal stresses and displacements were then compared to those
observed experimentally. A spring constant of 350 k/in. pro-
vided the best correlation between the analytical and experi-
mental stresses. The calibration weld size was
3

16
-in. The
spring constant for other weld sizes were determined assum-
ing a linear relationship between the shear stiffness of the
weld and the spring constant. All models contained the same
number of spring elements per linear inch of weld.
Table 2.
Experimental Specimen Details
Weld Configuration
W
1
a
W
2
a
W
3
a
Test No.
Test
Designation Member
A
g
b
(in.
2
)
Length
(in.)
Size
(in.)
Length
(in.)
Size
(in.)
Length
(in.)
Size
(in.)
1 P-L1-1a PL4
3

8
1.47 5
1

2
1

4
5
1

2
1

4
2 P-L1-1b PL3
1

4
0.785 4
1

4
1

4
4
1

4
1

4
3 P-L1-2 PL3
1

4
0.783 4
1

4
1

4
4
1

4
1

4
4 P-L1-3 PL3
1

4
0.781 4
1

4
1

4
4
1

4
1

4
5 P-L2-1 PL3
1

4
0.785 5
1

4
5
1

4
6 P-L2-2 PL3
1

4
0.784 5
1

4
5
1

4
7 P-L2-3 PL3
1

4
0.777 5
1

4
5
1

4
8 P-B-l PL3
1

4
0.780 3
1

4
3
1

4
3
1

4
9 P-B-2 PL3
1

4
0.777 3
1

4
3
1

4
3
1

4
10 P-B-3 PL3
1

4
0.783 3
1

4
3
1

4
3
1

4
11 L-L-1 L22
3

16
0.760 4
1

2
3

16
4
1

2
3

8
12 L-L-2 L22
3

16
0.761 4
1

2
3

16
4
1

2
3

8
13 L-L-3 L22
3

16
0.756 4
1

2
3

16
4
1

2
3

8
14 L-B-1a L43
1

4
1.68 3
1

2
1

4
4
1

4
3
1

2
1

4
15 L-B-1b L22
3

16
0.756 3
3

16
2
3

16
3
7

16
16 L-B-1c L22
3

16
0.771 3
3

16
2
3

16
3
7

16
17 L-B-2 L22
3

16
0.764 3
3

16
2
3

16
3
7

16
18 L-B-3 L22
3

16
0.750 3
3

16
2
3

16
3
7

16
19 L-T-1 L43
1

4
1.67 4
1

4

20 C-L-1 C34. 1 1.29 5
3

8
5
3

8
21 C-L-2 C34.1 1.28 5
3

8
S
3

8
22 C-L-3 C34.1 1.26 5
3

8
5
3

8
23 C-B-1 C34.1 1.24 5
3

16
3
3

16
5
3

16
24 C-B-2 C34.1 1.19 5
3

16
3
3

16
5
3

16
25 C-B-3 C34.1 1.22 5
3

16
3
3

16
5
3

16
26 C-T-1 C45.4 1.58 4
1

4

27 C-T-2 C34.1 1.19 3
3

16

a. See Figure 3.
b. Gross area based on measured cross section dimensions.
THIRD QUARTER / 1993 81
General Results
In all tests with cross section ruptures, the failure occurred
after the cross section yielded. The yielding was qualitatively
observed by flaking of whitewash and quantitatively ob-
served in the instrumented specimens from strain readings
and in all specimens from load cell readings, which exceeded
the yield load. Ideally, specimens used to determine shear lag
coefficients would rupture at the critical section prior to
yielding on the gross cross section. Shear lag coefficients
determined from tests in which yielding occurs on the gross
cross section prior to rupture on the net cross section may
differ from those determined from tests that do not yield prior
to rupture. This hypothesis has not been verified in the study
reported here, nor in past studies. As indicated in the review
of past research, this limitation was also present in most of
the tests conducted as part of the research reported by Ches-
son and Munse.
6
Yield lines, indicated by flaking of the whitewash, gener-
ally were not observed within the directly connected portion
of the members. In some instances, the portion of the cross
section that was not connected, e.g. outstanding angle leg,
showed indications of yielding. Yielding mostly occurred in
the portion of the member between the welded ends.
Experimental results are given in Table 3. The experimen-
tal shear lag coefficients, U
e
, were calculated as the ratio of
the failure load to the rupture strength (gross area tensile
stress). The shear lag coefficients for the specimens that did
not exhibit rupture at the critical cross section can be taken at
least equal to those shown in Table 3. The values for these
tests do not explicitly represent shear lag coefficients because
rupture was not the controlling limit state. Calculated shear
lag coefficients, U
t
, were determined using Equation 1, except
for the plate specimens and the transversely welded speci-
mens. Coefficients for the plate specimens were determined
from the current AISC specifications,
2,3
which give the coef-
ficients according to:
a. If l 2w . . . . . . . . U = 1.0
b. If 2w > l 1.5w . . . . U = 0.87
c. If 1.5w > l w . . . . U = 0.75
where
w = plate width (distance between welds)
The shear lag coefficient for longitudinally welded plates can
also be calculated using Equation 1 with each half of the plate
treated independently. Therefore, x
_
would be one-fourth of
the plate width. These values are not shown in Table 3.
Coefficients for the transversely welded members were cal-
culated as the ratio of the area of the directly connected
elements to the gross area. This is also an AISC specification
provision. The calculation procedure for the shear lag coeffi-
cients is deemed acceptable if the ratios of experimental to
calculated shear lag coefficients, given in Table 3, fall within
a 10 percent scatter band, i.e. 0.9 to 1.1. A similar evaluation
was made for bolted and riveted tests reported by Munse and
Chesson.
11
Plate Specimens
Results are summarized in Table 3 and as indicated, the plate
tests can be divided into three groups according to the speci-
fication shear lag coefficients of 0.75, 0.87, and 1.0. (Values
computed using Equation 1, as described in the previous
paragraph, are 0.82, 0.85, and 1.0.) Two of the groups have
only longitudinal welds and one has both transverse and
longitudinal welds.
The plate specimens exhibited tearing across the member
at the critical section, which was at the end of the welds.
Yielding in the plates was first observed at the critical cross
section at the end of the welds. None of the plate specimens
displayed significant out-of-plane effects. For all of the plate
specimens the ratio of U
e
/ U
t
was greater than 0.9. Six of the
nine tests have values of U
e
/ U
t
greater than or equal to 1.1.
Note that a failure load was not obtained for Test 1 because
the testing machine capacity was exceeded. However a shear
lag coefficient of 0.92 is reported. This represents the maxi-
mum load applied, and the true coefficient would have been
greater.
Longitudinal strains were recorded across the width of the
instrumented plate specimens near the end of the welds. The
strain gage locations for Test 3 are shown in Figure 5. Strains
were converted to stresses and distributions plotted along the
critical section. The stress distribution at various load levels
within the elastic range of material behavior for Test 3 is
illustrated in Figure 6.
Note the unsymmetrical stress distributions shown in Fig-
ure 6(a), most likely caused by imbalance in the longitudinal
welds or eccentrically applied load. The welds were detailed
for a balanced configuration, but given the stress distribu- Fig. 4. Finite element model of typical angle.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 82
tions, they apparently were not fabricated symmetrically. One
would expect the distributions to be symmetric if the welds
were balanced and the load applied concentrically.
Figure 7 is a plot of the stress distributions at approximately
the critical section for Tests 3, 6, and 8. The strain gages were
within 0.5 in. of the critical section. Note that both specimens
welded only longitudinally exhibited unsymmetric stress dis-
tributions, while the specimen that was welded with both
longitudinal and transverse welds exhibited essentially a sym-
metric distribution. Assuming that the stress distribution
would be symmetric if the welds were balanced, then the
experimental stresses may be modified to permit an evalu-
ation of the influence of the longitudinal weld length. Figure 8
is a plot of the distributions in which the symmetric strain
readings, e.g. gages 1 and 5 and gages 2 and 4, were averaged
prior to converting the values to stresses.
In Figure 8, the three tests show similar distribution pat-
terns, but with varying magnitudes of stress. The highest
stresses occurred for Test 8, which had 3-in. longitudinal
welds along with a transverse weld. Test 3, which had 4
1

4
-in.
longitudinal welds, exhibited lower stresses than Test 6,
which had 5-in. longitudinal welds. The variations in the
stresses for Tests 3 and 6 ranged between 3 and 7 percent,
while the stresses for Test 8 were 57 percent higher than
those for Test 6.
The analytical, based on finite element analyses, and ex-
Table 3.
Experimental Results
Test No.
Test
Designation
F
y
(ksi)
F
u
(ksi)
Failure Load
Per Member,
P
u
/ 2 (k)
Calculated
Shear Lag
Coefficient, U
e
Theoretical
Shear Lag
Coefficient, U
t
U
e
/ U
t
1 P-L1-1a 48.4 73.2 99.0
a
0.92 0.75 1.23
2 P-L1-1b 51.9 73.0 53.7 0.94 0.75 1.25
3 P-L1-2 51.9 73.0 56.0 0.98 0.75 1.31
4 P-L1-3 51.9 73.0 57.5 1.00 0.75 1.33
5 P-L2-1 51.9 73.0 55.9 0.98 0.87 1.13
6 P-L2-2 51.9 73.0 55.8 0.98 0.87 1.13
7 P-L2-3 51.9 73.0 54.4 0.96 0.87 1.10
8 P-B-1 51.9 73.0 51.2 0.90 1.00 0.9
9 P-B-2 51.9 73.0 56.1 0.99 1.00 0.99
10 P-B-3 51.9 73.0 55.7 0.97 1.00 0.97
11 L-L-1 54.1 81.1 50.0 0.81 0.87 0.93
12 L-L-2 54.1 81.1 50.5 0.82 0.87 0.94
13 L-L-3 54.1 81.1 50.4 0.82 0.87 0.94
14 L-B-1a 47.8 71.3 98.7 0.82 0.80 1.03
15 L-B-1b 54.1 81.1 49.5
b
0.81
16 L-B-1c 54.1 81.1 50.0 0.80 0.81 0.99
17 L-B-2 54.1 81.1 46.2 0.75 0.81 0.93
18 L-B-3 54.1 81.1 48.8 0.80 0.81 0.99
19 L T-1 47.8 71.3 55.8
b
0.59
20 C-L-1 57.0 75.5 87.0
c
0.89 0.91 0.98
21 C-L-2 57.0 75.5 86.7
c
0.90 0.91 0.99
22 C-L-3 57.0 75.5 86.9
c
0.91 0.91 1.00
23 C-B-1 55.7 76.6 85.1
c
0.92 0.91 1.01
24 C-B-2 55.7 76.6 84.0
c
0.92 0.91 1.01
25 C-B-3 56.0 77.1 83.1 0.88 0.91 0.97
26 C-T-1 58.5 77.6 60.0
b
0.44
27 C-T-2 51.1 73.8 32.3
b
0.49
a. Testing machine capacity exceeded.
b. Weld failure.
c. Gross cross section failure away from welds.
THIRD QUARTER / 1993 83
perimental stresses for Tests 3, 6, and 8 can be compared with
Figures 911. The trend for each of the three cases is similar.
The experimental stresses near the center of the plate are
approximately the same, or somewhat less, than the calcu-
lated stresses. Experimental stresses nearer the edge of the
plate are greater than the calculated stresses. This trend may
have been caused by the stopping or starting of the welding
process, causing imperfections at the critical section in the
form of gouges or notches due to blow out. These imperfec-
tions would result in stress concentrations adjacent to the edge
of the plate, which would in turn cause yielding earlier in the
loading process. Stress concentrations caused by the imper-
fections at the critical cross section were not considered in the
finite element model.
The experimental results for the plate specimens indicated
that the longitudinal weld length appears to not influence the
rupture strength based on shear lag effects. This observation
was reinforced by the elastic finite element results which
showed virtually the same stress distribution at the critical
section for models in which the weld length is 3, 4
1

4
and 5 in.
Neither the stress distribution at the critical section nor the
experimental shear lag coefficient were significantly affected
by the addition of the transverse weld, as compared to the
specimens with only longitudinal welds. However, note that
the differences in longitudinal weld length were relatively
small.
Angle Specimens
All but two of the angle specimens exhibited a tearing
failure, with the tearing initiating at the welded toe. The
welds sheared in Tests 15 and 19. The outstanding legs of
the specimens generally exhibited more signs of yielding
at the critical section, evident by whitewash flaking, than
the area of the angles directly connected to the gusset
Fig. 5. Strain gage locations for Test 3.
Fig. 6. Experimental stresses for Test 3.
Fig. 7. Experimental stresses at the critical section
for Tests 3, 6, and 8.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 84
plates. This behavior was observed in both longitudinally
welded specimens and specimens with a combination of
longitudinal and transverse welds, and was attributed to com-
bined stress caused by out-of-plane eccentricity. Yielding
generally was first visible at the heel of the angle in the
connected leg.
Significant out-of-plane bending occurred in the speci-
mens fabricated with 2L43
1

4
. Bending in the plane of the
connected leg also occurred in Test 14. The welds for this
specimen were not balanced, nor was the centerline of the
angle coincident with the centerline of the end plates, i.e. the
line of load application. Negligible out-of-plane bending was
observed in the specimens fabricated with 2L22
3

16
.
The ratios of U
e
/ U
t
, given in Table 3, for the angle speci-
mens vary from 0.93 to 1.03, the majority of the values being
less than or equal to 0.99. These results indicate that the
calculated shear lag coefficients compare well to the experi-
mental results for this group of tests. However, it is interesting
to note that the experimental values for all but one test were
between 0.80 and 0.82, while the calculated values ranged
between 0.80 and 0.87. The increased length of the welds for
Tests 1113 did not affect the shear lag coefficient as ex-
pected. As with the plate tests, the addition of the transverse
welds did not affect the maximum loads, or shear lag coeffi-
cients, for the angle tests.
Channel Specimens
The predominant limit state observed in the channel tests was
rupture in the cross section away from the welded region.
Each specimen in the series of longitudinally welded chan-
Fig. 8. Modified experimental stresses at the critical section
for Tests 3, 6, and 8.
Fig. 11. Analytical vs. experimental stresses for Test 8.
Fig. 10. Analytical vs. experimental stresses for Test 6.
Fig. 9. Analytical vs. experimental stresses for Test 3.
THIRD QUARTER / 1993 85
nels, and all but one of the longitudinally and transversely
welded channels, failed in the center of the specimen. This
was attributed to the combined state of stress induced in the
members by the out-of-plane load eccentricity.
Initial yielding was generally concentrated in the channel
flanges and near the web-flange intersection. In the welded
area, as with the angles, yielding was visible in the outstand-
ing flanges while none was present in the directly connected
web. The propagation of yielding into the channel flanges was
attributed to the combined axial stress and bending stress due
to out-of-plane eccentricity.
Note in Table 3 that all the experimental shear lag coeffi-
cients for the channels ranged between 0.88 and 0.92. How-
ever, the predominant limit state was rupture of the gross cross
section approximately halfway between the two ends of the
specimen, and not rupture of the net section. These results
agree with the observed practical upper limit of 0.9 that
Chesson and Munse
11
identified from their studies. Due to
eccentricities and fabrication imperfections in welded speci-
mens, an upper limit of 0.9 for the shear lag coefficient
appears prudent. As with the plate and angle specimens, the
maximum loads and experimental shear lag coefficients were
not affected by the addition of a transverse weld.
SPECIFICATION REVIEW AND
RECOMMENDED REVISIONS
The AISC design specifications for shear lag pertained only
to bolted or riveted connections prior to the inclusion of
welded members in the 1986 LRFD Specification,
2
and sub-
sequently in the 1989 ASD Specification.
3
Welded members
are treated similar to bolted members to maintain continuity
in the specifications. However, the provisions for welded
members are not clear in all instances and have therefore
raised questions regarding their application. A review of the
questions and related issues, along with recommended
changes to the specification, are presented in this section of
the paper.
Although the specification indicates that welded members
are subject to shear lag reductions, there is no minimum weld
length criterion to distinguish between different coefficient
values in Chapter B3. The first set of subparagraphs a, b, and
c in section B3 identify a minimum fastener length indirectly,
by specifying a minimum number of fasteners, in the direction
of stress for bolted or riveted connections but not for welded
connections. In fact, because welding is not mentioned in
subparagraphs a, b, or c, while bolting and riveting are, it is
unclear that the definitions apply to welded members. Nev-
ertheless, these sections are intended to be applicable to
welded specimens.
Another unclear portion of the specification pertains to the
use of members with both transverse and longitudinal welds.
A specification provision is given for members connected by
only transverse welds. If the first group of subparagraphs a,
b, and c are assumed to apply only to members with longitu-
dinal welds, then no provisions exist for cases in which a
combination of longitudinal and transverse welds are used.
Results of this study indicate that the addition of a transverse
weld does not significantly affect the rupture strength com-
pared to a specimen with only longitudinal welds.
The shear lag provision for members welded only with
transverse welds specifies that the effective area shall be the
area of the connected element. Reviewing the limit states of
weld shear and shear lag, summarized in the Appendix of this
paper, indicates that weld shear will always control the
strength if fillet welds are used. If partial- or full-penetration
welds are used then the present specification provision is
appropriate.
According to the specification commentary, previous re-
search
4
determined that plates welded only longitudinally can
fail prematurely due to shear lag if the distance between the
welds is too great. Thus, a minimum weld length equal to the
plate width or distance between the welds, w, is required.
Currently, the specification does not consider shear lag a
limiting factor as long as the weld length is greater than twice
the plate width. Two shear lag coefficients are specified for
intermediate ranges of longitudinal weld length between w
and 2w. Results from this study indicate that the weld lengths
greater than the distance between the welds have little influ-
ence on the shear lag coefficient. However, due to the limited
number of tests conducted and to the small size of the mem-
bers, no modifications are recommended to the shear lag
coefficients for longitudinally welded plates.
Reviewing the AWS
4
results, along with the statement by
Munse and Chesson
11
that efficiencies greater than 90 percent
are seldom observed, an upper limit for U of 0.9 is deemed
appropriate. This is also consistent with the upper limit that
appears in the current specifications
2,3
in section B3 subpara-
graph a. The strength of welded tension members is reduced
due to the coupled effects of shear lag, stress concentrations,
and eccentricities. The stress concentrations are due to the
sudden change in stiffness caused by the presence of the weld,
or to notches or gouges created at the critical section by the
welding process. Although all play a role in reducing the
strength, it is difficult to determine the relative participation
of each component. Using an empirical approach, such as the
shear lag coefficient, is an approximate way to account for all
the effects.
Recommended Revisions to the
Specification and Commentary
Recommended revisions to the specification were developed
jointly by the authors of this paper and the AISC Task Com-
mittee 108Connections and Force Introduction. The rec-
ommended changes address all of the issues identified in the
previous section. All of the recommended changes apply to
section B3 of the AISC Specifications.
2,3
Subparagraphs a, b, and c that follow the line Unless a
larger coefficient can be justified by tests or other rational
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 86
criteria.... should be replaced with a single equation for U,
given by:
U = 1
x
_
l
0.9 (3)
The specific values of U, given for certain groups of sections
in subparagraphs a, b, and c, are acceptable for use in lieu of
the values calculated from Equation 3 and may be retained in
the commentary for continued use by designers.
The section that addresses sections only connected with
transverse welds should be modified to include all shapes, not
just W, M, or S shapes and structural tees cut from these
shapes. A provision should be added to indicate that this
section is only applicable if partial- or full-penetration welds
are used, and is not applicable if fillet welds are used as the
transverse weld type.
In addition to changing the commentary to incorporate the
information of subparagraphs a, b, and c, several other
changes would help to clarify specification revisions. Several
of these are indicated in the following paragraph. Although
the primary focus of the research project reported in this paper
was welded tension members, literature and specification
provisions for bolted members were reviewed. The comments
made in the following paragraphs pertaining to bolted mem-
bers are the authors judgment based on that review.
For any given profile and connected elements, x
_
is a geo-
metric property. It is defined as the distance from the connec-
tion plane, or face of the member, to the member centroid, as
indicated in Figure 12. Note that the member may be a
portion of the cross section for particular cases. Connection
length, l, is dependent upon the number of fasteners or length
of weld required to develop the given tensile force, and this
in turn is dependent upon the mechanical properties of the
member and the strength of the fasteners or weld used. The
length l is defined as the distance, parallel to the line of force,
between the first and last fasteners in a line for bolted con-
nections. The number of bolts in a line, for the purpose of
determining l, is determined by the line with the maximum
number of bolts in the connection. For staggered bolts, use
the out-to-out dimension for l (See Figure 13). For welded
connections, l is the length of the member parallel to the line
of force that is welded. For combinations of longitudinal and
transverse welds (see Figure 14), l is the length of longitudinal
weld because the transverse weld does not significantly affect
the rupture strength based on shear lag. The presence of the
transverse weld does little to get the load into the unattached
portions of the member.
SUMMARY AND CONCLUSIONS
The purpose of this investigation was to review the shear lag
provisions for welded tension members relative to those for
bolted members, and to make recommendations for pertinent
specification changes. Experimentally, three different mem-
ber types and three different weld configurations were con-
sidered. Results of 27 tests were reported. Longitudinal
stresses were determined analytically in a finite element
study. Experimental strains were determined directly from
Fig. 12. Definition of x
_
for various members.
Fig. 13. Definition of l for bolted members with staggered holes.
Fig. 14. Definition of l for welded members.
THIRD QUARTER / 1993 87
tensile tests. The analytical and experimental stress patterns
in the elastic region were compared. Shear lag criteria are
recommended based on the experimental results. The current
AISC provisions have been reviewed and revisions recom-
mended.
The recommended revisions to the specification are based
on the results of the experimental and analytical studies
reported here, a review of the specification and judgment of
the authors. In particular the definitions of x
_
in Figure 14 and
l in Figures 15 and 16 are based on the authors judgment.
Further, the hypothesis that the net section failure is due to a
combination of the shear lag effect and stress concentration
caused by welding induced imperfections is also based on the
authors judgment, given the insight gained from the research
program. Each of these topics would require further study to
be proven explicitly.
The following conclusions have been drawn from the
experimental and analytical investigations:
1. Shear lag controlled the strength of the angle and plate
specimens.
2. For plates connected only by longitudinal welds, con-
nection length had little influence on the experimental
shear lag coefficient.
3. The transverse weld in the angle members welded both
longitudinally and transversely did not increase the
shear lag coefficient as expected. The experimental
shear lag coefficients of the longitudinally welded an-
gles and the angles with both longitudinal and transverse
welds were equivalent.
4. Shear lag will not control the strength of tension mem-
bers connected only by transverse fillet welds. Weld
shear will be the controlling limit state, regardless of
electrode strength or fillet weld size. This conclusion
does not apply to partial- or full-penetration welds.
5. Due to the small size of the experimental specimens in
this study as well as past studies, caution should be
exercised when applying the design provisions to much
larger tension members. There is a need for some limited
confirmatory testing on large tension members designed
so that shear lag effects control the strength.
6. The recommended upper limit for the shear lag coeffi-
cient is 0.9.
7. The implementation of the recommended changes to
AISC specifications and commentaries would result in
a simpler, more uniform approach to the application of
shear lag provisions to bolted and welded tension mem-
bers. The changes should result in fewer questions re-
garding the application of the provisions.
ACKNOWLEDGMENTS
Financial support for this project was provided by the Ameri-
can Institute of Steel Construction and Virginia Tech. Valu-
able assistance was provided throughout the project by Nestor
Iwankiw of AISC. Steel for the test specimens was provided
by Montague-Betts Co., Inc. The technical input provided by
the AISC Task Committee 108, particularly comments by J.
W. Fisher, T. M. Murray and W. A. Thornton, was very
beneficial to the authors. Additionally, comments provided by
D. R. Sherman proved very useful. The authors are grateful
to all of the above. Many of the specimens were tested in the
materials testing laboratory at Virginia Military Institute. The
willingness of C. D. Buckner and D. K. Jamison to permit
access to the laboratory made the completion of the experi-
mental parts of this project possible, and for this the authors
are grateful.
APPENDIX
The strength of members welded only with transverse fillet
welds will not be controlled by the rupture based on shear lag
effects, but rather will be controlled by weld shear. This is true
regardless of the steel or electrode strength. This can be
shown by considering the following parameters:
A
e
= bt = area of connected element
E70XX electrodes (fillet welds)
A36 steel
The strength of the tension member based on rupture is given
by:
P
n
= F
u
A
e
= 0.75(58 ksi)bt (A1)
where for tension rupture is 0.75.
The weld strength is given by:
R
w
= 0.6F
EXX
A
w
(A2)
where for weld shear is 0.75 and R
w
= nominal weld
resistance.
If the weld area, A
w
, is taken as (0.707t)b (the maximum
possible dimension for a fillet weld made along the edge of
plate element), then R
w
becomes
R
w
= 0.75(0.6)(70 ksi)(0.707t)b = 0.75(29.7 ksi)bt (A3)
Comparing Equations A1 and A3, one observes that the weld
strength, Equation A3, is less and will therefore control the
strength. The same conclusion will be reached for any prac-
tical combination of weld electrode and steel. If the sub-
merged arc process were used, the weld strength result, Equa-
tion A3, would increase, but it would still remain less than
Equation A1.
REFERENCES
1. American Institute of Steel Construction, Specification
for the Design, Fabrication, and Erection of Structural
Steel for Buildings with Commentary, Chicago: AISC,
1978.
2. American Institute of Steel Construction, Load and Re-
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 88
sistance Factor Design Specification for Structural Steel
Buildings, Chicago: AISC, 1986.
3. American Institute of Steel Construction, Specification
for Structural Steel Buildings, Allowable Stress Design
and Plastic Design, Chicago: AISC, 1989.
4. American Bureau of Welding, Report of Structural Steel
Welding Committee, American Welding Society, New
York, 1931.
5. Chesson, E., Jr. Behavior of Large Riveted and Bolted
Structural Connections. Thesis presented to the Univer-
sity of Illinois in partial fulfillment of the requirements
for the degree of Doctor of Philosophy. Urbana, 1959.
6. Chesson, E., Jr. and Munse, W. H. Riveted and Bolted
Joints: Truss-Type Tensile Connections, Journal of the
Structural Division, ASCE, Vol. 89(ST1), 1963, pp. 67
106.
7. Davis, R. P. and Boomslitter, G. P. Tensile Tests of
Welded and Riveted Structural Members, Journal of the
American Welding Society, 13(4), 1934, 2127.
8. Fisher, J. W., personal communication, ATLSS, Lehigh
University, Bethlehem, PA, 1990.
9. Gonzalez, L. and Easterling, W. S. Investigation of the
Shear Lag Coefficient for Welded Tension Members, Re-
port No. CE/VPI-ST 89/13. Virginia Polytechnic Institute
and State University, Blacksburg, VA, December, 1989.
10. Kulak, G. L., Fisher, J. W. and Struik, J. H. A., Guide to
Design Criteria for Bolted and Riveted Joints, 2nd Edi-
tion, John Wiley and Sons, New York, 1987.
11. Munse, W. H. and Chesson, E., Jr. Riveted and Bolted
Joints: Net Section Design, Journal of the Structural
Division, ASCE, Vol. 89(ST1), 1963, pp. 107126.
12. Swanson Analysis Systems, Inc., ANSYS Engineering
Analysis System Users Manual, Vol. I and II. Houston,
PA, 1989.
THIRD QUARTER / 1993 89
ABSTRACT
In this paper, a useful design aid for determining the values
of the initial connection stiffness R
ki
, the ultimate moment
capacity M
u
, and the shape parameter n of a three-parameter
power model describing the moment-rotation curve (M-
r
) of
semi-rigid connections with angles is prepared for its use in
the practical design of flexibly jointed frames with angles. A
set of nomographs allows the engineer to rapidly determine
the M-
r
curve for a given connection.
Applying the design aid, numerical simulations on drift and
column moment of a flexibly jointed frame with angles are
illustrated.
1. INTRODUCTION
The aim of the work described in this paper is to provide a
practical procedure for the analysis and design of semi-rigid
frames with angles. To this end, a set of nomographs allows
the engineer to determine rapidly the values of the initial
connection stiffness R
ki
, the ultimate moment capacity M
u
,
and the shape parameter n of a three-parameter power model
describing the M-
r
curve of connections. In this develop-
ment, a data base of steel beam-to-column connections was
first built and simple procedures to enable engineers to assess
this M-
r
behavior were then formulated. Using this data base,
extensive comparisons were made with the results of tests on
actual connections providing final confirmation of the valid-
ity of the three-parameter power model. The model is recom-
mended for general use in semi-rigid frame analysis.
In this paper, we have established a design procedure for
connections with angles. The three-parameter power model
is adopted to represent the nonlinear M-
r
curve proposed
previously by Richard and Abbott (1975). The three parame-
ters in this model are the initial connection stiffness R
ki
, the
ultimate moment capacity of connection M
u
, and the shape
parameter n. The values of R
ki
and M
u
can be determined by
a simple mechanical procedure with an assumed failure
mechanism (Kishi and Chen, 1990). Herein, we prepared a
useful design aid for the values of these parameters corre-
sponding to given angles and beam, or the main parameters
of connection angles for given values of R
ki
and M
u
. The shape
parameter n can be determined as a linear function of
log
10

0
(Kishi and Chen et al. 1991) which is an empirical
equation based on experimental data installed in the Program
SCDB (Chen and Kishi 1989), where
0
= M
u
/ R
ki
.
Using the nomographs for R
ki
and M
u
and the empirical
equation for n, we can determine rapidly the nonlinear M-
r
curve of connections with angles. Then, using the second-or-
der elastic analysis program FRAME formulated by Goto and
Chen (1987), we can analyze the flexibly jointed frame in a
simple manner (Chen and Lui, 1991). As an illustrative ex-
ample, studies of a four-bay, two-story frame with variable
beam section and/or length or thickness of connection angles
made using this analysis are presented.
2. ASSUMPTIONS AND NOTATIONS
In this paper, four types of connections with angles are
considered: single/double web-angle connections and top-
and seat-angle with/without double web-angle connections as
shown in Figures 1 to 3. To prepare the design charts for the
initial connection stiffness R
ki
and the ultimate moment ca-
pacity M
u
, the dimensions used for angle as shown in Figure 4
are defined as:
t = angle thickness
k = gauge distance from heel to the top of fillet
l = angle length
g = distance between heel to the center of fastener closest
to web or flange of beam
W = nut width
I
0
= t
3
/ 12 = geometrical moment of inertia
M
0 =
y
t
2
/ 4 = pure plastic bending moment
where
y
is the yielding stress of steel, and I
0
and M
0
the values
per unit length of plate element of angle. We assume that top
angle and seat angle have the same dimensions.
Design Aid of Semi-rigid Connections
for Frame Analysis
N. KISHI, W. F. CHEN, Y. GOTO, and K. G. MATSUOKA
N. Kishi is associate professor, civil engineering, Muroran
Institute of Technology, Muroran, Japan 050.
W. F. Chen is George E. Goodwin distinguished professor of
civil engineering and head of structural engineering, School of
Civil Engineering, Purdue University, West Lafayette, IN.
Y. Goto is professor, civil engineering, Nagoya Institute of
Technology, Nagoya, Japan 466.
K. G. Matsuoka is professor, civil engineering, Muroran Insti-
tute of Technology, Muroran, Japan 050.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 90
Furthermore, we shall introduce the following non-dimen-
sional parameters:

g
l
,
l
t
,
d
t
,
k
t
,
W
t
,
t
w
t
t
in which d is the height of beam and subscripts t and w
denote top angle and web angle respectively.
3. CHARTS FOR R
ki
and M
u
IN
CONNECTIONS WITH ANGLES
To prepare the charts for the initial connection stiffness R
ki
and the ultimate moment capacity M
u
, the equations devel-
oped previously by Kishi and Chen (1990) are used. All charts
are basically related to the parameter . Extensive compari-
sons of the analytical solutions with experimental test results
can be found in the paper by Kishi and Chen (1990).
3.1 Single/Double Web-Angle Connections
Using a simple mechanical procedure described in the paper
cited above, the values of R
kiw
and M
u
for single web-angle
connections are formulated as:
R
kiw
EI
0w

12 cos h(
w
)
7.8

'

(
w
) cos h (
w
) sin h (
w
)

;
)
(1)
M
uw
M
0w
t
w

(2
w
+ 1)
3

w
2
(2)
Fig. 1. Single web-angle connections.
Fig. 3. Top and seat angle with/without double web-angle connections.
(a) Top and seat angle with double web-angle connection (b) Top and seat angle without double web-angle connection
Fig. 2. Double web-angle connections.
THIRD QUARTER / 1993 91
in which = 4.2967 and
w
is defined as

w

w

1

w
+

w
2
_

,
(3)

w
in Equation 2 is obtained by solving Equation 4 which
is derived by combining the Drucker interaction equation
between bending moment and shearing force with the Tresca
yielding criterion as

w
4
+ (
w

w

w
)
w
1 0 (4)
The non-dimensional initial connection stiffness as a func-
tion of
w
for single web-angle connections is shown in
Figure 5. The non-dimensional ultimate connection moment
is shown in Figure 6 in which
w
is taken as abscissa and
w
is varied from 5 to 20 with an increment of 5 for
w
= 1.5
and 2.0.
The case of double web-angle connections can be obtained
simply by doubling the values found from these charts.
3.2 Top and Seat Angle Without Double
Web-Angle Connection
Assuming that the center of rotation is located at the angle leg
adjacent to the compression beam flange and the top angle acts
as a cantilever beam to resist surcharged moment, the initial
connection stiffness R
kits
is obtained as (Kishi and Chen, 1990).
R
kits
EI
0t
(1 +
t
)
2
D
ts
(5)
in which D
ts
is a function of
t
and
t
and those D
ts
and
t
are
given by
D
ts

3

t
(
t
2

2
+ 0.78)
(6)

t

t

(1 +
t
)
2
t
(7)
The ultimate moment capacity M
uts
is obtained by assuming
a simple failure mechanism. The equation for M
uts
is given by
M
uts
M
0t
t
t

t

'

1 +
t
(1 +
t

+ 2(
t
+
t
)

;
)
where the variable
t
is a non-dimensional ultimate shearing
force acting at the plastic hinge. Here, as in the case of single
Fig. 4. Main parameters for an angle.
Fig. 6. Ultimate moment capacity for
single web-angle connections.
Fig. 5. Initial connection stiffness for
single web-angle connections.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 92
web-angle connections, it is obtained by solving Equation 9
as

t
4
+
t

t
1 0 (9)
in which
t

is defined as

t

t
(10)
The distributions of the coefficient D
ts
for R
kits
with respect
to
t
are shown in Figure 7 in which
t
is varied from 5 to 40
with an increment of 5. Figure 8 is the chart for the non-
dimensional ultimate moment capacity M
uts
. Figure 8(a)
shows the results with the variation of
t
for
t
= 40 and 80,
while Figure 8(b) shows the variation of
t
(10 to 80) for
t
=
5 and 40.
3.3 Top and Seat Angle With Double
Web-Angle Connection
In this type of connection, the initial connection stiffness R
ki
and the ultimate moment capacity M
u
can be evaluated by
separating the top- and seat-angle part and the double web-
angle part as
R
ki
EI
0t

R
kits
EI
0t
+
R
kiw
EI
0t
,
M
u
M
0t
t
t

M
uts
M
0t
t
t
+
M
uw
M
0t
t
t
(11,12)
Since the top- and seat-angle part of the equations derived
above are also applicable for the case of the top- and seat-an-
gle connections, Figure 7 can be used for R
kits
/ EI
0t
and Figure
8 for M
uts
/ M
0t
t
t
in this type of connections.
As for the web angle, it acts as a cantilever beam similar to
the behavior of the top angle, the initial connection stiffness
R
kiw
is related to the double web-angle connection part as
(Kishi and Chen, 1990)
R
kiw
EI
0t
(1 +
t
)
2
D
w
(13)
in which D
w
is
D
w

3
2
w
(
w
2

2
+ 0.78)
(14)
where
w
is defined the same as
t
in Equation 7.
In the limit state, choosing a simple failure mechanism of
web angle and taking moment about the center of rotation at
the angle leg adjacent to the compression beam flange, the
ultimate moment capacity M
uw
is
M
uw
M
0t
t
t

w
(1 +
w
)

'

w
1
3(
w
+ 1)

w
+
w
+
1

;
)

3
(15)
in which
w
is obtained by solving Equation 4, since the
mechanism assumed here is the same as in the single web-an-
gle connections.
The distributions of the coefficient D
w
for R
kiw
/ EI
0t
are
given in Figure 9. M
uw
has a total of five variables:
w
,
w
,
w
,
, and
w
. Taking
w
as abscissa, two types of charts are
prepared in this study. In the first case,
w
is varied from 5 to
35 and/or 40 with an increment of 5 while the values of
w
,
, and
w
are kept constant (Figure 10). In the second case,

w
instead of
w
is varied in a similar manner (Figure 11).
Though it is easy to obtain these curves for arbitrary values
of these parameters, we consider here only two cases = 0.5
or 1.0 and
w
= 1.5 or 2.0.
4. DETERMINATION OF M-
r
CURVE
OF CONNECTION
It is a simple matter to obtain the values of R
ki
and M
u
for given
dimensions of angle or to determine the angle dimensions for
given values of R
ki
and M
u
. Moreover, we must also determine
the nonlinear characteristics of connection behavior for a
structural analysis (Chen, 1987). The three-parameter power
model is adopted here to represent these characteristics of
semi-rigid connections which is a simplification of the four-
parameter power model proposed previously by Richard and
Abbott (1975).
Assuming m = M / M
u
,
r
/
0
and
0
M
u
/ R
ki
and
introducing the shape parameter n, the power model used here
has the simple form
m

(1 +
n
)
1
n
(16)
Figure 12 shows the M-
r
curves of a connection with
several values of shape parameter n. In one extreme, if the
shape parameter n is taken to be infinity, the model reduces
to a bilinear curve with the initial connection stiffness R
ki
and
Fig. 7. Coefficient D
ts
for R
kits
of top and seat angle
without double web-angle connections.
THIRD QUARTER / 1993 93
Fig. 8. Ultimate moment capacity for top and seat angle without double web-angle connections.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 94
the ultimate moment capacity M
u
. The principal merit of the
present model is a significant saving of computing time in a
non-linear structural analysis program, since the connection
moment M can be represented as a function of relative rota-
tion
r
. Furthermore, the tangent connection stiffness R
k
can
be directly obtained without iteration.
As for the shape parameter n, we use the following proce-
dure for its determination (Kishi and Chen et al. 1991):
1. The shape parameter n for each experimental test data is
numerically determined first by the least mean square
technique of the test data with Equation 16.
2. The shape parameter is assumed to be a linear function
of log
10

0
. Using a statistical technique for n values
obtained from the above procedure, empirical equation
for n for each connection type are determined.
Figure 13 shows comparisons of the distributions of n
values of the empirical equation with the experimental test
data installed in the program SCDB (Kishi and Chen, 1989).
From these numerical considerations, we conclude that
within the current practice of the range of the connection
variables, the three-parameter power model with the shape
parameter n obtained from the empirical equation can be
applied in practical design (Kishi and Chen et al. 1991). In
this study, we set the shape parameter n to be constant for the
region of
0
less than the smallest one obtained from experi-
mental test data. The equation refined for each connection
type is listed in Table 1.
5. NUMERICAL EXAMPLE OF STRUCTURAL
ANALYSIS OF FLEXIBLY JOINTED FRAME
In this study, a four-bay, two-story frame used by Lindsey
(1987) is taken as the frame of basic skeleton for the present
numerical analysis (Figure 14).
W824 and W831 sections are used as the external and
internal columns respectively and the frames are placed 25 ft
center to center. The loads are: floor dead load: 65 psf, roof
dead load: 20 psf, reduced floor live load: 40 psf, and roof
live load: 12 psf. Wind loads are assumed to be 15 psf with a
shape factor of 1.3. Two types of load combination are con-
sidered referring to AISC-LRFD specification (1986). One is
the unfactored loads (D+L+W) to check the drift under service
load. Another is the factored loads (1.2D + 0.5L + 1.3W) to
check the frame stability. Load intensities are W
R
= 0.80 k/ft,
W
F
= 2.70 k/ft, P
R
= 2.925 kip, P
F
= 6.581 kip for the unfactored
loads and W
R
= 0.75 k/ft, W
F
= 2.54 k/ft, P
R
= 3.803 kip, P
F
=
8.556 kip for the factored loads.
In the present study, we adopt the top and seat angle with
double web-angle connections as a part of the beam-to-col-
umn connection. The combinations of beam, column, and
connection angle for several cases are listed in Table 2 in
which the size of the columns is constant for each case. Beams
used in Cases 1 and 2 are stronger than those in Cases 3 and 4.
Fig. 9. Coefficient D
w
for R
kiw
/ EI
0t
of top and seat angle
with double web-angle connections.
Table 1.
Empirical Equations for Shape Parameter n
Type No. Connection Type n
I Single web-angle connection 0.520 log
10

0
+ 2.291
0.695
log10 0 > 3.073
3.073
II Double web-anlge connection 1.322 log
10

0
+ 3.952
0.573
log10 0 > 2.582
2.582
III Top- and seat-angle connection
(with double web angle)
1.398 log
10

0
+ 4.631
0.827
log10 0 > 2.721
2.721
IV Top- and seat-angle connection
(with double web angle)
2.003 log
10

0
+ 6.070
0.302
log10 0 > 2.880
2.880
THIRD QUARTER / 1993 95
Fig. 10. Ultimate moment for the variation of
w
for top and seat angle with double web-angle connections.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 96
Fig. 11. Ultimate moment for the variation of
w
for top and seat angle with double web-angle connections.
THIRD QUARTER / 1993 97
Heavy hex structural bolts with 1-in. nominal size are used as
fasteners for all cases.
6. NUMERICAL RESULTS
6.1 Characteristics of M-
r
Curve of Connections
The M-
r
curves of connections used are shown in Figure 15
in which the dimensions of beams and angles are specified in
Case 1 and the length of top and seat angles is six inches.
6.2 Drift of Frame in Case Surcharging Unfactored Loads
The general configurations of deformation of flexibly jointed
frame under the service loads are shown in Figure 16 com-
paring with the results of rigid connections, in which the
sections of members are taken as l
t
= 6 in. and t
t
=
1

4
-in. and/or
1

2
-in. as in Cases 2 and 4. Though R
ki
and M
u
in the case of
t
t
=
1

2
-in. may be twice than that of t
t
=
1

4
-in. as we can see in
Figure 15, the drift of roof in the case of t
t
=
1

4
-in. is less than
twice that of t
t
=
1

2
-in. The drifts for t
t
=
1

2
-in. and
1

4
-in. are
almost two to three times than that of the result of rigid
connections, respectively.
Distributions of the non-dimensional roof drift ( / H) for
each case are shown in Figure 17 taking l
t
or g
t
as abscissa in
which and H are roof drift and height of frame respectively.
From this figure, we can select some dimensions for top and
seat angles for a given drift. For example, if the maximum
drift is set to be / H =
1

300
, we can choose three types of
angles to meet this requirement as
For Cases 1 and 2:
t
t
=
1

2
-in. g
t
= 2.75 in. l
t
= 6 in.
and
t
t
=
1

2
-in. g
t
= 2.50 in. l
t
= 5 in.
For Cases 3 and 4:
t
t
=
1

2
-in. g
t
= 2.50 in. l
t
= 6 in.
6.3 Frame Stability in Case Surcharging Factored Loads
Bending moment diagrams of the frame in Cases 2 and 4 with
l
t
= 6 in. and t
t
=
1

4
-in. and/or
1

2
-in. under the factored loads
are shown in Figure 18 together with the results of rigid
connections. The bending moments on floor beams show a
large difference between the semi-rigid and rigid connections.
On the other hand, the differences on other members are
smaller than those of the floor beams.
The non-dimensional end moments of columns of Cases 1
to 4 are tabulated in Table 3 together with the results of rigid
connections and the B
1
, B
2
method as given in AISC-LRFD
specification (Chen and Lui, 1987). The reference values in
each case are obtained from a first order elastic analysis with
rigid connections. In these tables, all values at the fixed points
of flexibly jointed frame in all cases are greater than the Fig. 12. M-
r
curves for the three-parameter power model.
Table 2.
Combinations of Beam, Column, and Connection Angles
Beam and Column Sizes:
Case 1, 2 Case 3, 4
Floor beam W1850 W1846
Roof Beam W1422 W1219
External column W824 W824
Internal column W831 W831
Top- and Seat-Angle Sizes:
Size l g l g
T & S Angles L43
1

2
t
t
6 in. var. var. 2.5 in.
Web Angles L32
1

4
8.5 in. for floor
5.5 in. for roof
1.75 in. 8.5 in. for floor
5.5 in. for roof
1.75 in.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 98
Fig. 13. Comparison of the shape parameter n of the empirical equation with experimental test data.
THIRD QUARTER / 1993 99
reference value, and its maximum value is almost 3.5 times
greater than the reference value. Alternatively, almost all of
the bending moments at the top of the 1st floor columns are
less than the reference ones and the values at Node No. 2 of
Element No. 1 in Figure 14 for some cases have an opposite
sign for the reference one. The results of the flexibly jointed
frame in all cases are substantially different from the results
(1, 2, and 3) of rigid connections.
On the other hand, referring to the results of rigid connec-
tions, it is clear that the values (1) obtained from the second
order elastic analysis are almost the same with the values (2)
obtained from the B
1
, B
2
method with Equation H1-5 in the
AISC LRFD specification (1986).
7. SUMMARY AND CONCLUSIONS
A considerable amount of test data on semi-rigid connections
with angles has been collected and analyzed and simple
models developed in the past years. Against the background
Fig. 14. General view of a four-bay, two-story frame.
Fig. 15. M-
r
curves of connections in Case 1.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 100
of this information, design aids for a three-parameter connec-
tion model put forward recently by Kishi and Chen is pre-
pared here. A set of nomographs allows the engineer to rapidly
determine the values of initial connection stiffness R
ki
and the
ultimate moment capacity M
u
for a given connection, or the
basic dimensions of angles for given values of R
ki
and M
u
. The
general validity of the design procedures based on these
developed design charts is demonstrated by comparisons with
computed displacements and moments at service loads and
factored loads of a four-bay, two-story frame with semi-rigid
connections using a second-order elastic analysis program.
With the aid of these design charts, the present analysis
procedure for the design of semi-rigid frames with angles has
achieved both simplicity in use and, as far as possible, a
realistic representation of actual behavior. Taking this point
in conjunction with the demonstrated validity of the ap-
proach, it is recommended for general use.
REFERENCES
1. American Institute of Steel Construction, Load and Re-
sistance Factor Design Specification for Structural
Buildings, Chicago, IL, 1986.
2. Chen, W. F., editor, Joint Flexibility in Steel Frames,
Journal of Construction Steel Research, Special Issue,
Vol. 8, Elsevier Applied Science, London, 290 pp, 1987.
3. Chen, W. F. and Lui, E. M., Structural Stability: Theory
and Implementation, Elsevier, New York, 486 pp, 1987.
4. Chen, W. F. and Lui, E. M., Stability Design of Steel
Frames, CRC Press, Boca Raton, Florida, 380 pp, 1991.
5. Chen, W. F. and Kishi, N., Semi-Rigid Steel Beam-to-
Column Connections: Data Base and Modeling, ASCE
Journal Structural Engineering, 115(ST1), pp. 105119,
1989.
6. Goto, Y. and Chen, W. F, On the Computer-Based Design
Analysis for Flexibly Jointed Frames, Journal of Con-
struction Steel Research, 8, pp. 203231, 1987.
7. Kishi, N. and Chen, W. F., Data Based of Steel Beam-to-
Column Connections, Structural Engineering Report
No. CE-STR-80-26, School of Civil Engineering, Purdue
University, West Lafayette, IN, 653 pp, 1986.
8. Kishi, N. and Chen, W. F., Moment-Rotation Relations
of Semi-Rigid Connections with Angles, ASCE Journal
Structural Engineering, 116(ST7), pp. 18131834, 1990.
9. Kishi, N., Chen, W. F., Goto, Y. and Matsuoka, K. G.
Applicability of Three-Parameter Power Model to Struc-
tural Analysis of Flexibly Jointed Frames, ASCE Me-
chanics Computing in 1990s and Beyond, pp. 238242,
1991.
10. Lindsey, S. D. Design of Frames with PR Connections,
Journal of Construction Steel Research, 8, pp. 251260,
1987.
11. Richard, R. M. and Abbott, B. J., Versatile Elastic-Plastic
Stress-Strain Formula, ASCE Journal Engineering Me-
chanical Division, Vol. 101, No. EM4, pp. 511515, 1975.
NOMENCLATURE
d beam depth
D
ts
defined in Equation 6
D
w
defined in Equation 14
g distance between heel to the center of fastener
closest to web or flange of beam (Figure 4)
I
o
t
3
/ 12 = moment of inertia per unit length
k gauge distance from heel to the top of fillet (Figure
4)
l length of the angle
M connection moment
M
0
y
t
2
/ 4 = plastic bending moment capacity per unit
length
m M / M
u
M
u
ultimate connection moment capacity
n shape parameter of the three-parameter power
model defined in Equation 16
R
k
tangent connection stiffness
R
ki
initial connection stiffness
V shear force
V
o
y
t / 2 = plastic shear capacity per unit length
t angle thickness (Figure 4), subscripts w and t may
be used to refer to web angle and top angle
respectively
W nut diameter (Figure 4) Fig. 16. General deformations of frame under service loads.
THIRD QUARTER / 1993 101
Fig. 17. Distributions of roof drift under service loads.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 102

r
/
o

r
relative rotation of connection

o
M
u
/ R
ki

w
V / V
o
= non-dimensional ultimate shear force
parameter in web angle

t
non-dimensional ultimate shear force parameter at
the plastic hinge

y
yield stress of steel

g / l

w
defined in Equation 3

defined in Equation 10

t
defined in Equation 7
l / t
d / t
k / t
W / t
t
w
/ t
t
Fig. 18. Bending moment diagrams of frame
under factored loads.
THIRD QUARTER / 1993 103
Table 3.
Non-dimensional End Moments of Columns
g
t
= 2.5 in. of Case 1
T&S Angle with W Angle Connection Rigid Connection
Reference
Value
(kip-in.)
Elment
No.
Node
No.
t
t

1

4
-in. t
t

5

16
-in. t
t

3

8
-in. t
t

1

2
-in.
(1)
Exact
B
1
, B
2
Method
(2)
Equation
H1-5
(3)
Equation
H1-6
1 1 3.00 2.74 2.44 1.91 1.16 1.22 1.33 85.28
2 1.37 1.10 0.73 0.12 0.48 0.46 0.21 32.46
2 4 1.39 1.33 1.26 1.15 1.06 1.08 1.12 282.48
5 0.61 0.64 0.67 0.73 1.06 1.07 1.11 297.07
3 7 1.54 1.47 1.39 1.27 1.07 1.07 1.12 254.32
8 0.74 0.77 0.82 0.88 1.07 1.07 1.12 242.04
4 10 1.68 1.61 1.53 1.39 1.07 1.07 1.12 230.75
11 0.91 0.96 1.01 1.09 1.10 1.07 1.13 195.40
5 13 1.18 1.17 1.18 1.19 1.05 1.03 1.06 290.13
14 0.60 0.68 0.77 0.95 1.03 1.02 1.04 379.62
6 2 0.50 0.62 0.75 0.95 1.00 1.00 1.00 301.11
3 0.40 0.51 0.62 0.81 1.00 1.00 1.00 252.37
7 5 0.26 0.20 0.16 0.18 1.01 1.01 1.01 146.61
6 0.84 0.89 0.93 0.97 1.02 1.01 1.01 125.05
8 8 0.60 0.44 0.32 0.29 1.01 1.02 1.01 63.59
9 1.56 1.64 1.68 1.61 1.02 1.02 1.02 66.26
9 11 3.95 2.84 1.87 1.07 0.90 0.88 0.88 9.81
12 8.43 8.74 8.77 7.67 1,10 1.09 1.10 12.30
10 14 0.54 0.63 0.73 0.92 1.00 1.00 1.00 367.51
15 0.73 0.84 0.95 1.07 1.00 1.00 1.00 329.61
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 104
Table 3 (cont.).
Non-dimensional End Moments of Columns
t
t
=
3

8
-in. of Case 2
T&S Angle with W Angle Connection Rigid Connection
Reference
Value
(kip-in.)
Elment
No.
Node
No.
l
t
= 4.0 in. l
t
= 4.5 in. l
t
= 5.0 in. l
t
= 5.5 in. l
t
= 6.0 in. l
t
= 6.5 in.
(1)
Exact
B
1
, B
2
Method
(2)
Equation
H1-5
(3)
Equation
H1-6
1 1 2.73 2.65 2.58 2.51 2.44 2.38 1.16 1.22 1.33 85.28
2 1.07 0.99 0.90 0.81 0.73 0.64 0.48 0.46 0.21 32.46
2 4 1.33 1.31 1.29 1.28 1.26 1.25 1.06 1.08 1.12 282.48
5 0.64 0.65 0.65 0.66 0.67 0.68 1.06 1.07 1.11 297.07
3 7 1.46 1.44 1.43 1.41 1.39 1.38 1.07 1.07 1.12 254.32
8 0.78 0.79 0.80 0.81 0.82 0.82 1.07 1.07 1.12 242.04
4 10 1.61 1.58 1.56 1.55 1.53 1.51 1.07 1.07 1.12 230.75
11 0.96 0.97 0.99 1.00 1.01 1.02 1.10 1.07 1.13 195.40
5 13 1.18 1.18 1.17 1.17 1.18 1.18 1.05 1.03 1.06 290.13
14 0.68 0.70 0.73 0.75 0.77 0.79 1.03 1.02 1.04 379.62
6 2 0.63 0.67 0.70 0.73 0.75 0.78 1.00 1.00 1.00 301.11
3 0.52 0.55 0.58 0.60 0.62 0.65 1.00 1.00 1.00 252.37
7 5 0.19 0.18 0.17 0.16 0.16 0.15 1.01 1.01 1.01 146.61
6 0.89 0.90 0.91 0.92 0.93 0.94 1.02 1.01 1.01 125.05
8 8 0.42 0.39 0.36 0.33 0.32 0.30 1.01 1.02 1.01 63.59
9 1.64 1.66 1.67 1.68 1.68 1.69 1.02 1.02 1.02 66.26
9 11 2.74 2.48 2.24 2.04 1.87 1.73 0.90 0.88 0.88 9.81
12 8.75 8.79 8.80 8.80 8.77 8.72 1.10 1.09 1.10 12.30
10 14 0.63 0.66 0.68 0.70 0.73 0.75 1.00 1.00 1.00 367.51
15 0.85 0.88 0.91 0.93 0.95 0.97 1.00 1.00 1.00 329.61
THIRD QUARTER / 1993 105
Table 3 (cont.).
Non dimensional End Moments of Columns
g
t
= 2.5 in. of Case 3
T&S Angle with W Angle Connection Rigid Connection
Reference
Value
(kip-in.)
Elment
No.
Node
No.
t
t

1

4
-in. t
t

5

16
-in. t
t

3

8
-in. t
t

1

2
-in.
(1)
Exact
B
1
, B
2
Method
(2)
Equation
H1-5
(3)
Equation
H1-6
1 1 3.49 3.16 2.79 2.10 1.18 1.25 1.37 78.31
2 0.66 0.49 0.25 0.34 0.64 0.63 0.45 48.26
2 4 1.46 1.39 1.31 1.17 1.06 1.08 1.13 284.67
5 0.55 0.59 0.63 0.70 1.06 1.07 1.11 299.00
3 7 1.62 1.54 1.45 1.29 1.07 1.07 1.12 255.26
8 0.68 0.73 0.78 0.86 1.07 1.07 1.12 241.40
4 10 1.78 1.69 1.59 1.42 1.07 1.07 1.12 230.72
11 0.85 0.91 0.97 1.08 1.10 1.07 1.14 192.74
5 13 1.21 1.20 1.19 1.19 1.05 1.03 1.06 297.85
14 0.55 0.63 0.73 0.93 1.03 1.02 1.04 392.94
6 2 0.39 0.53 0.68 0.92 1.00 1.00 1.00 330.36
3 0.44 0.54 0.65 0.82 1.00 1.00 1.00 291.10
7 5 0.46 0.36 0.27 0.22 1.02 1.01 1.01 151.84
6 0.65 0.74 0.84 0.97 1.02 1.01 1.00 121.47
8 8 1.02 0.79 0.56 0.37 1.02 1.02 1.02 66.24
9 1.20 1.36 1.51 1.61 1.03 1.02 1.02 64.67
9 11 7.00 5.37 3.66 1.67 0.92 0.88 0.90 9.54
12 6.40 7.18 7.83 7.58 1.08 1.08 1.07 12.11
10 14 0.56 0.64 0.73 0.91 1.00 1.00 1.00 398.34
15 0.63 0.75 0.88 1.03 1.00 1.00 1.00 363.95
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 106
Table 3 (cont.).
Non dimensional End Moments of Columns
t
t
=
3

8
-in. of Case 4
T&S Angle with W Angle Connection Rigid Connection
Reference
Value
(kip-in.)
Elment
No.
Node
No.
l
t
= 4.0 in. l
t
= 4.5 in. l
t
= 5.0 in. l
t
= 5.5 in. l
t
= 6.0 in. l
t
= 6.5 in.
(1)
Exact
B
1
, B
2
Method
(2)
Equation
H1-5
(3)
Equation
H1-6
1 1 3.15 3.06 2.96 2.88 2.79 2.71 1.18 1.25 1.37 78.31
2 0.47 0.42 0.36 0.30 0.25 0.19 0.64 0.63 0.45 48.26
2 4 1.39 1.37 1.34 1.33 1.31 1.29 1.06 1.08 1.13 284.67
5 0.59 0.60 0.61 0.62 0.63 0.64 1.06 1.07 1.11 299.00
3 7 1.53 1.51 1.49 1.47 1.45 1.43 1.07 1.07 1.12 255.26
8 0.73 0.74 0.75 0.77 0.78 0.79 1.07 1.07 1.12 241.40
4 10 1.69 1.66 1.64 1.62 1.59 1.57 1.07 1.07 1.12 230.72
11 0.91 0.93 0.94 0.96 0.97 0.99 1.10 1.07 1.14 192.74
5 13 1.20 1.19 1.19 1.19 1.19 1.19 1.05 1.03 1.06 297.85
14 0.64 0.66 0.69 0.71 0.73 0.75 1.03 1.02 1.04 392.94
6 2 0.54 0.58 0.61 0.65 0.68 0.71 1.00 1.00 1.00 330.36
3 0.55 0.57 0.60 0.62 0.65 0.67 1.00 1.00 1.00 291.10
7 5 0.35 0.33 0.30 0.28 0.27 0.25 1.02 1.01 1.01 151.84
6 0.75 0.77 0.79 0.82 0.84 0.85 1.02 1.01 1.00 121.47
8 8 0.77 0.71 0.66 0.61 0.56 0.52 1.02 1.02 1.02 66.24
9 1.37 1.41 1.45 1.48 1.51 1.53 1.03 1.02 1.02 64.67
9 11 5.24 4.80 4.38 4.00 3.66 3.35 0.92 0.88 0.90 9.54
12 7.22 7.42 7.58 7.72 7.83 7.91 1.08 1.08 1.07 12.11
10 14 0.64 0.67 0.69 0.71 0.73 0.75 1.00 1.00 1.00 398.34
15 0.77 0.80 0.83 0.86 0.88 0.91 1.00 1.00 1.00 363.95
THIRD QUARTER / 1993 107
Steel tubes of relatively thin wall thickness filled with high-
strength concrete have been used in building construction in
the U.S. and Far East Asian countries. This structural system
allows the designer to maintain manageable column sizes
while obtaining increased stiffness and ductility for wind and
seismic loads. Column shapes can take the form of tubes or
pipes as required by architectural restrictions. Additionally,
shop fabrication of steel shapes helps insure quality control.
In this type of construction, in general, at each floor level
heavy steel beam is framed to these composite columns.
Often, these connections are required to develop shear yield
and plastic moment capacity of the beam simultaneously.
This paper summarizes results and recommendations from
a pilot study conducted to develop a moment-resisting steel
connection detail for connecting steel beams to composite
columns of the type described above. The focus of this pilot
study was on composite columns having a square or rectan-
gular cross section.
CURRENT PRACTICE
Beam-column connections in concrete-filled steel tubes are
usually constructed by directly welding the steel beam to the
tube when connections are required to develop plastic mo-
ment capacity of the beam. Current design practices for these
connections rely heavily on the judgment and experience of
individual designers, with little research and testing informa-
tion available.
When beams are welded or attached to steel tubes through
connection elements, complicated stiffener assemblies are
required in the joint area within the column. However, weld-
ing of the steel beam or connecting element directly to the
steel tube of composite columns should be avoided for the
following reasons:
1. Transfer of tensile forces to the steel tube can result in
separation of the tube from the concrete core, thereby
overstressing the steel tube. In addition, the deformation
of the steel tube will increase connection rotation, de-
creasing its stiffness.
2. Welding of the thin steel tube results in large residual
stresses because of the restraint provided by other con-
nection elements.
3. The steel tube is designed primarily to provide lateral
confinement for the concrete which could be compromised
by the additional stress due to the welded connection.
POSSIBLE CONNECTION DETAIL
With these considerations in mind, attempts should be made
to prevent direct transfer of beam forces to the steel tube. Two
general types of connection details were envisioned, types A
and B.
Type A Connection Detail
Figure 1 shows one alternative in which forces are transmitted
to the core concrete via anchor bolts connecting the steel
elements to the steel tube. In this alternative, all elements
could be pre-connected to the steel tube in the shop. The nut
inside the steel tube is designed to accomplish this task. The
capacity of this type of connection would be limited with the
pull-out capacity of the anchor bolts and local capacity of the
tube.
Another variation of the same idea is shown in Figure 2,
where connecting elements would be embedded in the core
concrete via slots cut in the steel tube. In this variation slots
must be welded to connection elements after beam assembly
for concrete confinement. The ultimate capacity of this detail
also would be limited to the pull-out capacity of the connec-
tion elements and the concrete in the tube.
Type B Connection Detail
Another option is to pass the beam completely through the
column (see Figure 3). This type of connection is believed to
be the most suitable. In this type of detail a certain height of
column tube, together with a short beam stub passing through
the column and welded to the tube, could be shop fabricated
to form a tree column. The beam portion of the tree
column could then be bolted to girders in the field. A com-
bination of analytical and experimental investigations was
undertaken to comprehend and identify the force transfer
Atorod Azizinamini is an assistant professor, Department of Civil
Engineering, University of Nebraska-Lincoln, Lincoln, NE.
Bangalore Prakash, structural engineer with Nabih Youssef
and Associates, Los Angeles, California, formerly a graduate
student, Department of Civil Engineering, University of Ne-
braska, Lincoln, NE.
A Tentative Design Guideline for a New Steel Beam
Connection Detail to Composite Tube Columns
ATOROD AZIZINAMINI and BANGALORE PRAKASH
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 108
mechanism and suggest a tentative design procedure for this
type of connection.
ANALYTICAL INVESTIGATION
To investigate the performance of the connection detail in
which the beam completely passes through the column (here-
after referred to as a through connection), detailed finite
element analyses were conducted. The finite element model
used in these analyses consisted of a three-dimensional model
of the column with a small portion of the beam extending from
the column. In these analyses concrete cracking and non-lin-
ear behavior of the steel elements were modeled. In addition,
the interface between steel and concrete elements was care-
fully modeled.
Results of the analyses were used to identify the force
transfer mechanism between the steel beam and composite
column in the joint region, and to identify the effects of some
of the connection details on its performance. Major conclu-
sions from the analytical investigation associated with the
through beam connection detail are discussed in the following
section.
Figure 4 shows the force transfer mechanism observed
from the analyses. The portion of the steel tube between the
beam flanges acts as a stiffener, resulting in a concrete com-
pression strut which assists the beam web within the joint in
carrying shear. The effectiveness of the compression strut was
shown to be increased to a limit by increasing the thickness
of the steel plate between the beam flanges. The width of the
concrete compression strut on each side of the beam web in
the direction normal to the beam web is approximately equal
to half the beam flange width.
A compressive force block is created when beam flanges
are compressed against the upper and lower columns (Figure
4). The width of this compression block is approximately
equal to the width of the beam flange. In the upper and lower
columns shown in Figure 4 the compressive force, C, is shown
to be balanced by the tensile force provided by an embedded
rod in the concrete and possibly welded to the beam flanges.
This rod was not modeled in the finite element model, forcing
the steel liner plate to carry this tensile force.
Since one of the objectives of this phase of the study is to
devise means to improve connection performance, it is bene-
ficial to require rods be attached to beam flanges as shown in
Figure 4. The presence of such rods is believed to make the
beam web within the joints stiffer and reduce the stress level
in the steel tube.
EXPERIMENTAL INVESTIGATION
To gain additional insight of the behavior of the through beam
connection detail, one test specimen representing approxi-
mately a one-half scale model of a prototype column used in
high-rise building construction in seismic zones was con-
Fig. 1. Type A connection detail using anchor bolt. Fig. 2. Type A connection detail using embedded elements.
THIRD QUARTER / 1993 109
structed and tested. The prototype column consists of a 4-ft
(1.22-m) square hollow tube with a 2-in. (50.8-mm) wall
thickness, 10-9 (3.28-m) story height, and W30x99 beam
section framing to the column. In this particular building the
W30x99 beams were welded directly to the steel tube. To
prevent overstressing of the steel tube a complicated scheme
of stiffener assemblies was placed inside the hollow tube
directly behind the beam section.
Figure 5 shows the general configuration of the test speci-
men. The height of the column from the beams top and
bottom flanges to the support point is 31
11

16
inches (0.8 m)
and represents the distance from the floor to the inflection
point in the upper and lower stories of a building frame
subjected to lateral loading (assuming the inflection point to
be located at mid-height of the column). The length of the
beam extending from each side of the column is 27 inches
(0.69 m). This length was selected such that the beams
cross-section shear yield and plastic moment capacities
would develop simultaneously.
Figure 6 shows the different components of the test speci-
men. The test specimen consisted of three major components:
a. hollow steel tube made of A36 steel
b. hybrid built-up beam section
c. four #11 grade 60 reinforcing bars with anchor plates
welded to each end of the reinforcing bars
The hollow steel tube is 24 inches (0.6 m) square with
1

2
-in.
(12.7-mm) wall thickness. A half-scale model of the prototype
column (which has a 2-in. (50.8-mm) wall thickness) would
have required using 1-in. (25.4-mm) wall thickness in the test
specimen. However, only
1

2
-in. (12.7-mm) wall thickness is
used.
As shown in Figure 6, two slots in the shape of the beam
cross section were prepared on two faces of the steel tube.
These slots were used to pass the beam through the column.
Four holes were drilled on each flange of the beam within
the column as shown in Figure 6. These holes were used to
pass four #11 grade 60 (414 MPa) reinforcing bars through
the beam flanges. Reinforcing bars were then welded to the
beam flanges. As discussed earlier, these reinforcing bars
were provided to resist tensile forces in the lower and upper
columns arising from applied beam loads. The 42l-in.
(10250.825.4-mm) plates welded to each end of the rein-
forcing rods were intended to reduce the amount of slip in the
rebars. Excessive slip of the rebars could transfer large
tensile forces to the steel tube. It may be possible to achieve
this same objective by using longer rebars (develop the re-
bars) or by using a hook at the end of the rebars, particularly
since it has been reported that the use of steel plates at the end
of anchor bolts could potentially reduce their capacity.
1
The specimen was cast and cured in the vertical position.
The concrete compressive strength at time of testing was
14,000 psi (99 MPa).
TEST RESULTS
In this section the general behavior of the test specimen in
terms of function of the beam web within the joint are
described briefly. Further details are given elsewhere.
2
Figure 7 shows the location and orientation of six gages
attached to the beam web within the column. Also shown in
this figure is the direction of the applied beam loads. Data
from these gages, as shown in Figure 8, indicate that the beam
web within the joint is subjected primarily to compressive and
tensile strains along the lines GG and HH, respectively. This
Fig. 3. Through connection detail.
Fig. 4. Force transfer mechanism for through
beam connection detail.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 110
type of deformation indicates that the beam web experiences
shear type deformation.
Closer examination of data from gages shown in Figure 7
indicate that tensile strains along lines parallel to HH are
significantly larger than compressive strains parallel to line
GG. This observation can be explained as follows. The type
of shear deformation imposed on the beam web within the
joint results in the creation of a concrete compressive strut
parallel to line GG in Figure 7. This compressive strut acts as
a stiffener along the diagonal GG, consequently reducing the
compressive strain in the beam web in that direction. How-
ever, in the other direction (along line HH) tensile strains in
the web increase since concrete is not effective. This obser-
vation verified the force transfer mechanism deduced from
the analytical investigation and explained earlier.
BEHAVIORAL MODEL
Based on results of the finite-element analysis and experi-
mental results, a behavioral model in the form of equations
relating the applied external forces to the connections inter-
nal forces was developed. These equations are then used to
suggest a tentative design criteria for through-beam connec-
tion detail.
In developing the behavioral model the following assump-
tions were made:
a. Externally applied shear forces and moments at the joints
are known.
b. Failure is defined as the point at which the beam web
within the joint reaches its shear stress limit when exter-
nally applied forces are at their ultimate values.
c. At failure the concrete stress distribution is linear and
maximum concrete compressive stress is below its lim-
iting value.
The joint forces implied in assumption (a.) could be ob-
tained from analysis and requires the knowledge of applied
shear and moment at the joint at failure. These quantities are
assumed to be related as follows:
V
c = V
b
M
b
= l
1
V
b
M
c
= l
2
V
c
where V
b
and M
b
are ultimate beam shear and moment, respec-
tively, while V
c
and M
c
are ultimate column shear and moment,
respectively. Figure 9 shows these forces for an isolated
portion of a structure subjected to lateral loads.
The validity of assumption (c.) above could be justified for
the following reasons:
1. Column sizes for the type of construction considered in
this paper are generally much larger than the beam sizes.
2. The concrete type used in these columns is generally
high-strength concrete with compressive strength well
above 10,000 psi. The uniaxial stress-strain charac-
teristics of high-strength concrete exhibit a linear behav-
ior up to maximum strength, followed by a sharp de-
scending portion.
Derivation of Behavioral Model
The type of joint is shown in Figure 9. Figure 10 shows the
Free Body Diagram (FBD) of the beam web within the joint
Fig. 5. General configuration of test specimen.
Fig. 6. Different components of the test specimen:
a) hollow steel tube, b) hybrid built-up beam section,
c) four #11 reinforcing bars with anchor plates
welded to each end.
THIRD QUARTER / 1993 111
and upper column at ultimate load. With reference to Figure
10, the following additional assumptions are made in deriving
the Behavioral Model:
1. The concrete stress distribution is assumed to be linear.
The width of the concrete stress block is assumed to
equal b
f
, beam flange width.
2. As shown in Figure 10, strain distribution over the upper
column is assumed to be linear.
3. The steel tube and concrete act compositely.
4. The portion of the upper column shear, V
c
, transferred to
the steel beam is assumed to be C
c
, where C
c
is the
resultant concrete compressive force bearing against the
beam flange and is the coefficient of friction.
5. Applied beam moments are resolved into couples con-
centrated at beam flanges.
6. Resultant of concrete compression strut is along a diago-
nal as shown in Figure 10.
Considering the above assumptions and strain distribution
shown for the upper column in Figure 10, strain for different
connection elements could be related to
1
, steel tube strain
in tension.

c

a
d
c
a

1
(1)

sc

a d
1
d
c
a

1
(2)

st

d
c
d
1
a
d
c
a

1
(3)
where

c
= maximum compressive strain in steel tube and
concrete in compression

sc
= strain in steel rod in compression

st
= strain in steel rod in tension
Next, maximum stress in concrete and stresses in the steel
rod and steel tube could be calculated as follows:
f
c
E
c

c
(4)
f
sc
E
s

sc
(5)
f
lc
E
s

c
(6)
f
st
E
s

st
(7)
f
lt
E
s

1
(8)
where f
c
, f
sc
, f
lc
, f
st
, and f
lt
are maximum concrete concrete
compressive stress, stress in rod in compression, stress in steel
tube in compression, stress in rod in tension, and stress in steel
tube in tension, respectively.
Substituting Equations 1 through 3 in Equations 4 through
8 and multiplying Equations 4 through 8 by corresponding
area, the resultant forces for different connection elements
could be calculated as follows:
C
c

1

2
b
f

a
2
d
c
a
f
y1
(9)
C
1
b
f
t
1

a
d
c
a
f
y1
(10)
C
s
A
s

a d
1
d
c
a
f
y1
(11)
T
s
A
s

d
c
d
1
a
d
c
a
f
y1
(12)
T
1
b
f
t
1
f
y1
(13)
Using the FBD of the upper column shown in Figure 10,
Fig. 7. Location and orientation of gages
attached to beam web within the column.
Fig. 8. Strain data from gages attached to
beam web within the column.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 112
Equations 9 through 13, and satisfying vertical force equilib-
rium, the following equation could be obtained.
A
s

1
d
c
2a
[
1

2
b
f
a
2
A
1
(d
c
2a)] (14)
where
b
f
= beam flange width
d
c
= depth of the column
a = depth of the concrete compression block

= ratio of modulus of elasticity for concrete over


modulus of elasticity of steel
A
1
= effective area of steel tube = 2b
f
t
1
A
s
= area of steel rod at each corner of the beam
t
1
= thickness of steel tube
In defining A
1
it is assumed that a steel tube width equal to
two times the beam flange width is effective in carrying
tension and compression. This value was estimated from
experimental results.
Next, considering the moment equilibrium of the FBD of
the upper column shown in Figure 10 the following expres-
sion can be derived.

a
3
_

,
A
1
ad
c
+ A
s
(ad
c
2d
1
d
c
+ 2d
1
2
) +
1
1
]

2
b
f
a
2

d
c

a
3
_

,

1
1
]

f
y1
l
2
(d
c
a)
V
b
(15)
where
d
1
= distance between steel rod and steel tube
f
y1
= yield strength of steel tube
In Equation 15 f
y1
is the stress level the steel tube is
allowed to approach at ultimate condition. f
y1
could also be
viewed as the portion of the steel tube strength utilized to
resist the forces transferred by the connection. Based on the
limited experimental data obtained from this investigation it
is suggested that a value of 0.35 be used for .
Equations 14 and 15 relate the externally applied force,
V
b
, directly and the externally applied forces V
c
and M
c
indi-
rectly (through the coefficients and l
2
) to different connec-
tion parameters such as A
s
, A
1
, and a.
DESIGN APPROACH
Before proceeding with the steps necessary in designing the
through-beam connection detail, additional equations will be
derived to relate the shear stress in the beam web within the
joint to the compressive force in the concrete compression
strut and externally applied forces.
Considering the FBD of a portion of the beam web within
Fig. 9. Assumed forces on an interior joint
in a frame subjected to lateral loads.
Fig. 10. FBD of the upper column and beam
web within the joint area.
THIRD QUARTER / 1993 113
the joint area as shown in Figure 11 and satisfying the hori-
zontal force equilibrium, the following equation could be
derived:
V
w
+ C
st
cos + C
c

2M
b
d
b
0 (16)
where
V
w
= shear force in the beam web at ultimate condition
= arctan
d
b
d
c
Equations 14, 15, and 16 could be used to proportion the
through-beam connection detail.
Until further research is conducted the following steps are
suggested for designing the through-beam connection detail
following the LRFD format.
Step 1. From analysis, obtain factored joint forces.
Step 2. Select the following quantities: t
1
, b
f
, d
b
, d
c
, d
1
, f
y1
Step 3. Solving Equations 14 and 15 simultaneously, obtain
A
s
and a. This could be achieved using the trial and error
approach.
Step 4. Check stress in different connection elements.
Step 5. Assume the beam web yields at ultimate load. With
this assumption V
w
could be calculated as follows:
V
w
0.6F
yw
t
w
d
c
(17)
where
F
yw
= beam web yield stress
t
w
= thickness of the beam web
Step 6. Using Equation 16 calculate C
st
, compressive force in
the concrete compressive strut, and applied shear force to
concrete in the joint area.
Step 7. Check shear stress in concrete in the joint area. The
limiting shear force could be assumed to be as suggested by
ACI 352 [2]:
V
u
Rf
c
A
e
(18)
where
= 0.85
R = 20, 15, and 12 for interior, exterior, and corner joints,
respectively
f
c
= concrete compressive strength
It is suggested that the value of f
c
be limited to 100 psi,
implying that in the case of 15,000 psi concrete, for instance,
f
c
be taken as 100 rather than 122 as would be obtained from
V
u
calculations.
Until further research is conducted it is suggested that A
e
be calculated as follows:
A
e
2b
f
d
c
DESIGN EXAMPLE
Design a through-beam connection detail with the following
geometry and properties.
Given (Steps 1 and 2):
t
1
= 0.5 in.
b
f
= 5.5 in.
d
b
= 14.5 in.
d
c
= 24 in.
d
l
= 3.5in.
f
y1
= 36 ksi
F
yw
= 36 ksi
t
w
= 0.25 in.
= 0.85
l
2
= 32 in.
V
b
= 79 kips
M
b
= 1,660 in-kips
= 0.5
= 0.35
= 0.23
A
1
= 5.5 in
2
f
c
= 14 ksi
E
s
= 29,000 ksi (modulus of elasticity of steel)
E
c
= 6,670 ksi (modulus of elasticity of concrete)
Step 3: Using the trial and error approach and Equations 14
Fig. 11. FBD of the portion of the web within the joint area.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 114
and 15, calculate a and A
s
. For the first trial assume a = 8.5.
Equation 14 will result in:
A
s
=
1
24 2 8.5
[
1

2
(0.23)(5.5)(8.5)
2
5.5(24 2 8.5)
A
s
= 1.03 in
2
Substitute A
s
= 1.03 in.
2
in Equation 15 and calculate V
b
. If
the result is approximately equal to 79 kips the assumed value
of a is o.k. Equation 15 yields:
V
B
= [5.5 8.5 24 + 1.03(8.5 24 2 3.5 24 +
2 3.5
2
) +
1

2
(0.23)(5.5)(8.5)
2
(24

8.5 / 3)]
0.35 36
0.85 32(24 8.5)
V
b
= 64.3 kips 79 kips
Assume a = 9 inches. This will yield A
s
= 3.04 in.
2
, V
b
=
77
k
79
k
o.k.
Therefore, a = 9 inches and A
s
= 3.04 in.
2
Use two #11 Grade 60 deformed reinforcing bars. A
s
=
3.12 in.
2
Step 4: Check stresses in different connection elements
against their limit values. First calculate tensile strain in the
steel tube.

1
f
y1
/ E
s
0.35 36/29,000 0.000434 in./in.
Using Equations 1 and 4 calculate f
c
.
f
c
= 1.74 ksi < f
c
= 14 ksi o.k.
Using Equations 2, 3, 5, 6, 7, and 8 calculate stresses in
other connection elements. This yields:
f
sc
= 4.61 ksi <
c
F
y
0.85 60 51 ksi o.k.
f
1c
= 7.55 ksi <
c
F
y
0.85 36 30.6 ksi o.k.
f
st
= 9.65 ksi <
t
F
y
0.9 60 54 ksi o.k.
f
lt
= 12.6 ksi <
t
F
y
0.9 36 32.4 ksi o.k.
Step 5: Using Equation 17 calculate shear force in the beam
web:
V
w
= 0.6 36 0.25 24 = 129.6 kips
Step 6: Using Equation 16 calculate compressive force in
concrete compression strut.

= arctan 14.5/24 = 31.1


C
c =
1

2
b
f
(a
2
/ d
c
a) f
y1
C
c =
1

2
(0.23)(0.35)(5.5)(9
2
/ 24 9) 36 43 kips
V
w
+ C
st
cos() + C
c
(2M
b
/ d
b
) 0
129.6 + C
st
cos(31.1) + 0.5(43) (2 1,660) / 14.5 0
C
st
= 90.9 kips
Step 7: The shear force carried by concrete within the joint
between the beam flanges is assumed to be the horizontal
component, C
st
.
V
c
= C
st
cos()
V
c
= 90.9 cos(31.1) = 77.8
k
For the interior joint the shear capacity is
V
u
= (20)f
c
(2b
f
)(d
c
)
V
u
= 0.85(20)100[(25.5)(24)]/1,000 = 449
k
>
77.8
k
o.k.
SUMMARY AND CONCLUSIONS
The use of composite columns of the type described in this
paper is proven to be economical. This paper has summarized
a suggested connection detail (a through-beam connection
detail) for connecting steel beams to these columns as well as
tentative design guidelines. The information presented in this
paper is based on a pilot study and, therefore, it is suggested
that this information be viewed as a general guideline until
further research is carried out. It should also be noted that the
effect of axial load in the column on performance of the
connection was not considered. The intent of the paper is to
suggest an economical connection detail and outline a proce-
dure to comprehend its behavior through the behavioral
model presented.
ACKNOWLEDGMENT
The authors greatly appreciate the support provided by the
American Institute of Steel Construction which provided
partial funding of a graduate student. Valmont Industries of
Omaha, Nebraska is greatly appreciated for constructing the
test specimen. Special thanks are also due the Center for
Infrastructure Research at the University of Nebraska-
Lincoln for supporting this research. The authors thank Dr. J.
P. Colaco of CBM Engineers, Inc., Houston, Texas for his
helpful suggestions and great encouragement while pursuing
this research.
REFERENCES
1. Shipp, G. John and Haninger, R. Edward, Design of
Headed Anchor Bolts, AISC Engineering Journal, Sec-
ond Quarter, 1983, Vol. 20, No. 2 pp. 5869.
2. Prakash, A. Bangalore, Development of Connection De-
tail for Connecting Steel Beams to Composite Columns,
M.S. Thesis, Civil Engineering Department, University of
Nebraska-Lincoln, 1992.
3. ACI-ASCE Committee 352, Recommendations for De-
sign of Beam-Column Joints in Monolithic Reinforced
Concrete Structures, ACI Journal, Proceedings, Vol. 82,
No. 3, MayJune 1985, pp. 266283.
THIRD QUARTER / 1993 115
Eng ine e ring
J ourna l
AMERICAN INSTITUTE OF STEEL CONSTRUCTION, INC.
Pa g e 117: D. E. Alle n a nd T. M. Murra y
De sig n Crite rion for Vib ra tions Due to Wa lking
Pa g e 130: Won-Sun King a nd Wa i-Fa h Che n
LRFD Ana lysis for Se mi-Rig id Fra me De sig n
Pa g e 141: La i-Choon Ting , Na nd iva ra m E.
Sha nmug a m, a nd Se ng -Lip Le e
De sig n of I-Be a m to Box-Column Conne c tion Stiffe ne d
Exte r na lly
Pa g e 150: Disc ussionThoma s Sp uto a nd Ne stor R. Iwa nkiw
De sig n of Pip e Column Ba se Pla te s Und e r
Gra vity Loa d Thoma s Sp uto
Pa g e 152: Corre c tionASD/ LRFD Volume II
Conne c tions
She a r Ta b De sig n Ta b le s
Pa g e 153: 1993 Annua l Ind e x
4th Qua rte r 1993/ Volume 30, No. 4
ABSTRACT
A design criterion for walking vibrations of broader appli-
cation than previous criteria is proposed for steel floor or
footbridge structures. The criterion is based on the dynamic
response of steel structures to walking forces, as well as the
sensitivity of occupants to vibration motion. The criterion is
applicable to structures with natural frequencies below 9 Hz,
where resonance can occur with a harmonic of the step
frequency, but is extended beyond 9 Hz where footstep im-
pulse response becomes important.
INTRODUCTION
Walking, good for body and soul, sometimes produces vibra-
tions which are annoying to others. This is not a new problem.
Tredgold (1828) wrote that girders over long spans should be
made deep to avoid the inconvenience of not being able to
move on the floor without shaking everything in the room.
It also became common practice for soldiers to break step
when marching across bridges to avoid large and potentially
dangerous resonance vibrations. Both stiffness and resonance
are therefore important considerations in the design of steel
floor structures and footbridges for walking vibrations.
Stiffness has been taken into account for many years in the
design of floor structures using criteria dating from
Tredgolds time. A traditional stiffness criterion for residential
floors is to limit the deflection under 2 kPa (42 psf) to less
than span/360. This criterion is restricted to traditional wood
floor construction with high transverse stiffness. The Ameri-
can Institute of Steel Construction Allowable Stress Design
Specification (AISC, 1989) limits the live load deflection of
beams and girders supporting plastered ceilings to
span/360, a limitation which has also been widely applied to
steel floor systems in an attempt to control vibrations. A better
stiffness criterion applicable to all floor construction is to
limit the deflection due to 1 kN (225 lb.) concentrated load to
less than approximately 1 mm (0.04 in.).
Resonance, however, has been ignored in the design of
floors and footbridges for walking vibrations until recently.
Approximately 30 years ago, problems arose with walking
vibrations for steel-joist floors that satisfied code stiffness
criteria. Lenzen (1966) determined that damping and mass,
not stiffness, were the most important factors in preventing
unacceptable walking vibrations for these floors. To take
damping and mass into account, a simple dynamic design
criterion based on heel-impact response was developed (Al-
len and Rainer, 1976) and introduced 18 years ago into an
Appendix to the Canadian design standard for steel structures
(Canadian Standards Association, 1989), In 1981, Murray
recommended a similar dynamic design criterion based on
data from 91 floor measurements (Murray, 1981). More re-
cently a design criterion for footbridges has been introduced
into British and Canadian bridge standards based on reso-
nance response to a sinusoidal force (BSI, 1978; OHBDC,
1983).
Since these criteria were introduced, more has been learned
about the loading function due to walking, in particular that
resonance can occur at a harmonic multiple of the step fre-
quency. This has been verified by reviewing past cases of
vibration problems with steel joist and beam floors, most of
which corresponded to third harmonic resonance of the step
frequency (6 Hz floors approximately), but more recently also
to second harmonic resonance (4 Hz floors approximately).
Also the Canadian CSA criterion has recently been found not
to correctly predict the vibration behavior of two-way joist
girder systems.
In this paper a simple yet rational design criterion for
walking vibration is proposed based on harmonic resonance.
The criterion is calibrated to floor experience. It is similar to
one recently recommended by Wyatt (1989). The criterion is
extended to floor frequencies beyond 9 Hz to control impulse
vibration from footsteps.
VIBRATION LIMIT STATE
ACCELERATION LIMITS
International Standards Association (ISO, 1989; ISO, 1992)
recommends vibration limits below which the probability of
adverse reaction is low. Limits for different occupancies are
given in terms of rms acceleration as a multiple of the baseline
curve shown in Figure 1. For offices, ISO recommends a
D. E. Allen is senior research officer, Institute for Research in
Construction, National Research Council Canada, Ottawa, On-
tario, Canada.
T. M. Murray is Montague-Betts Professor of Structural Steel
Design, The Charles E. Via Department of Civil Engineering,
Virginia Polytechnic Institute and State University, Blacksburg,
VA.
Design Criterion for Vibrations Due to Walking
D. E. ALLEN AND T. M. MURRAY
FOURTH QUARTER / 1993 117
multiplier of 4 for continuous or intermittent vibrations and
60 to 128 for transient vibrations. Intermittent vibration is
defined as a string of vibration incidents such as those caused
by a pile driver, whereas transient vibration is caused rarely,
for example by blasting. Walking vibration is intermittent in
nature but not as frequent and repetitive as vibration caused
by a pile driver. It is therefore estimated that the multiplier for
walking vibration in offices is in the range of 5 to 8, which
corresponds to an rms acceleration in the range 0.25 to 0.4
percent g for the critical frequency range 4 to 8 Hz shown in
Figure 1. Based on an estimated ratio of peak to rms accel-
eration of approximately 1.7 for typical walking vibration, the
annoyance criterion for peak acceleration is estimated to be
in the range 0.4 to 0.7 percent g. From experience (Allen and
Rainer, 1976), a value of 0.5 percent g is recommended for
the frequency range 48 Hz. The resulting acceleration limit
for offices is shown in Figure 1.
For footbridges, ISO (1992) recommends a multiplier of
60 which, combined with an estimated ratio of peak to rms
acceleration of 1.7, results in a criterion of approximately ten
times the vibration limit for offices. People in shopping
centres will accept something in between, depending on
whether they are standing or sitting down. Suggested peak
acceleration limits for these occupancies are given in
Figure 1.
LOADING FUNCTION
Walking across a floor or footbridge produces a moving
repetitive force. Figure 2 shows the dynamic reaction at
mid-support of a footbridge due to a person walking across
it: the Fourier spectrum of the reaction clearly indicates the
presence of sinusoidal loading components at the first, sec-
ond, and third harmonic multiples of the step frequency. The
force, F, can therefore be represented in time by a Fourier
series
F = P (1 +
i
cos 2ift) (1)
where P is the persons weight, taken as 0.7 kN (160 lbs) for
design, f the step frequency, i the harmonic multiple,
i
is a
dynamic coefficient for the harmonic, and t is time. Table 1
recommends design values for these parameters based on test
information on dynamic coefficient (Rainer, et al, 1988) and
Fig. 1. Recommended acceleration limits
for walking vibration (vertical).
Fig. 2. Center support reaction produced by walking along a
footbridge on three supports (Rainer, et al, 1988).
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 118
observations of step frequencies which are in the range 1.9
0.3 Hz for offices.
Jogging, or more than one person walking in step, is a more
severe dynamic loading, but only for the first two harmonics.
Generally such cases are rare enough not to be a problem in
practice. Similarly a large group of people walking in an area
produces a greater dynamic loading at the step frequency
(2 Hz approximately), but lack of coherence at the higher
harmonics plus the damping effect of people has meant that,
except for footbridges close to entertainment events (Bach-
mann, 1992) such loading has not been a problem in practice.
RESPONSE
Walking across a footbridge or floor causes a complex dy-
namic response, involving different natural modes of vibra-
tion, as well as motion due to time variation of static deflec-
tion. The problem can be simplified by considering a person
stepping up and down at mid-span of a simply supported
beam which has only one mode of vibrationthe fundamen-
tal mode. Maximum dynamic response will occur when the
natural frequency corresponds to one of the harmonic forcing
frequencies. The steady-state acceleration, a, due to harmonic
resonance is given by (Rainer, et al, 1988),
a
g
=

i
P
0.5W

R
2
cos 2ift =
R
i
P
W
cos 2ift (2)
where W is the weight of the beam, is the damping ratio, g
is the acceleration due to gravity, and R is a reduction factor
discussed later. The factor 1 / (2) is the familiar dynamic
amplification factor for steady-state resonance and 0.5W / g
is the mass of an SDOF oscillator which is dynamically
equivalent to the simply supported beam of weight W vibrat-
ing in its fundamental mode. The other harmonics will also
produce steady-state vibrations at their forcing frequencies,
but the level of vibration is generally much smaller. For floor
structures, an exception occurs when there is resonance of two
modes of vibration at two multiples of the step frequency;
floor experience indicates, however, that only one resonant
mode whose frequency is near to the fundamental frequency
need be considered for design.
The reduction factor R is introduced into Equation 2 to take
into account (a) that full steady-state resonance is not
achieved when someone steps along the beam instead of up
and down at mid-span and (b) that the walker and the person
annoyed are not simultaneously at the location of maximum
modal displacement. Figure 3 shows test results for a person
walking across two simply supported footbridges which ver-
ify the harmonic resonance response model, Equation 2. The
value R = 0.56 in Figure 3a was determined by dynamic
analysis of a person walking across the footbridge (Rainer, et
al, 1988). It is recommended that for design R be taken as 0.7
for footbridges and 0.5 for floor structures having two-way
modal configurations.
PROPOSED DESIGN CRITERIA
Equation 2 predicts peak acceleration due to harmonic reso-
nance, R
i
P / W, which can be compared to the acceleration
limit, a
o
/ g shown in Figure 1. It is useful to express this in
terms of a minimum value of damping ratio times equivalent
mass weight (W):
W
R
i
P
a
o
/ g
(3)
Table 2 contains specific minimum values of W for the
values of dynamic loading (
i
P) from Table 1, acceleration
limit (a
o
/ g) from Figure 1 and reduction factor (R) recom-
mended above.
As shown in Figure 4 the results of Table 2 can be approxi-
mated by the following criterion for walking vibrations:
W K exp (0.35f
o
) (4a)
where f
o
is the fundamental natural frequency (Hz) and K is
a constant given in Table 3 which depends on the acceleration
limit for the occupancy. Equation 4a can be inverted to
express the criterion for walking vibrations in terms of mini-
mum fundamental natural frequency:
f
o
2.86 ln

K
W

(4b)
The following section provides guidance for estimating the
required floor properties for application of Equations 4.
DAMPING RATIO
The damping ratio depends primarily on non-structural com-
ponents and furnishings. The Canadian steel structures speci-
fication (CSA, 1989) recommends damping ratios of 0.03 for
a bare floor; 0.06 for a finished floor with ceiling, ducts,
flooring, and furniture; and 0.12 for a finished floor with
partitions. Murray (1991) recommends damping ratios of
0.01 to 0.03 for a bare floor, 0.01 to 0.03 for ceilings, 0.01 to
0.10 for mechanical ducts, and 0.10 to 0.20 for partitions.
These damping ratios, however, are based on vibration decay
resulting from heel impact and include a component for
Table 1.
Loading Function for Walking (See Equation 1)
Harmonic
i
Frequency Range
i f
Dynamic Load Factor

i
1 1.6 to 2.2 0.5
2 3.2 to 4.4 0.2
3 4.8 to 6.6 0.1
4 6.4 to 8.8 0.05
FOURTH QUARTER / 1993 119
geometric dispersion of vibration as well as frictional and
material damping. More recent testing of modal damping
ratios shows that the frictional and material damping ratios
are approximately half of the values determined from heel
impact tests. Based on available information (Wyatt, 1989;
ISO, 1992), Table 3 recommends damping values for use in
the proposed criterion, Equation 4.
NATURAL FREQUENCY, f
o
, AND EQUIVALENT
MASS WEIGHT, W
In the case of a simply supported panel such as a footbridge,
the natural frequency is equal to the fundamental beam fre-
quency of the panel and the equivalent mass weight is equal
to the panel weight. Floors of steel construction, however, are
two-way systems with many vibration modes having closely
spaced frequencies. Natural frequency and equivalent mass
weight of a critical mode in resonance with a harmonic of step
frequency is therefore difficult to assess. A dynamic modal
analysis of the floor structure can be used to determine the
critical modal properties, but there are factors that are difficult
to incorporate in the structural model. Composite action and
discontinuity conditions are two such factors, but more diffi-
cult to assess is the effect of partitions and other non-structural
components. An unfinished floor with uniform bays can have
a variety of modal pattern configurations extending over the
whole floor area, but partitions and other non-structural com-
ponents tend to constrain the modal configurations to local
areas in such a way that the floor vibrates locally like a single
two-way panel. The following simplified procedure is recom-
mended to estimate the properties of such a panel. Some of
the recommendations are based on judgment guided by floor
test experience. Further research is needed to obtain better
estimates, particularly for W.
The floor is assumed to consist of a concrete slab (or deck)
supported on steel joists or beams (open-web or rolled sec-
tions) which, in turn are supported on walls or on steel girders
between columns. The fundamental natural frequency, f
o
, and
equivalent mass weight, W, for a critical mode is estimated
by first considering a joist panel mode and a girder panel
mode separately and then combining them. If the joist span is
less than half the girder span, however, both the joist panel
mode and the combined mode should be checked against the
criterion, Equations 4.
Table 2.
Minimum Value of W determined from Equation 3
for Satisfactory Performance
Floor
Frequency
f
o
(Hz)
Office
Floors
kN (kips)
Shopping
Malls
kN (kips)
Footbridges
kN (kips)
1.6 to 2.2 28 (6.3) 9.3 (2.10) 4 (0.50)
3.2 to 4.4 14 (3.2) 4.7 (1.05) 2 (0.45)
4.8 to 6.6 7 (1.6) 2.3 ( 0.52) 1 (0.22)
6.4 to 8.8 3.5 (0.8) 1.1 (0.26) 0.5 (0.11)
Fig. 3. Peak response of two footbridge spans to a person walking across at different step frequencies (Rainer, et al, 1988).
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 120
In the following, the concrete modulus of elasticity is
assumed equal to 1.35 times that assumed in current structural
standards, the increase being due to the greater stiffness of
concrete under dynamic, as compared to static, loading. Also
for determining composite moment of inertia, the width of
concrete slab is equal to the member spacing but not more
than 0.4 times the member span. For edge members, it is half
of this value plus the projection of the slab beyond the
member center line.
Also the floor weight per unit area, w, should include the
sustained component of live load (approximately 0.5 kPa
(11 psf) for offices).
JOIST PANEL MODE
The joist panel mode is associated with the natural frequency
of the joist or beam alone. The natural frequency of this mode
can be estimated from the simple beam formula
f
j
= 0.18 g /
j
(5)
where
j
is the deflection of a beam or joist relative to its
supports due to the weight supported by the individual beam
or joist. Composite action is normally assumed provided the
joists are directly connected to the concrete slab by welds to
steel deck. Normally the joists or beams are assumed to be
simply supported unless dynamic restraint is verified by a
dynamic analysis or experiment. For open-web joists, shear
deformations should be included in the calculations for
j
.
The mass weight of the joist panel mode can be estimated
from
W
j
= wB
j
L
j
(6)
where w is the floor weight per unit area, L
j
the joist or beam
span, and B
j
the effective joist panel width determined from
B
j
= 2(D
s
/ D
j
)
1
4
L
j
(7)
where D
j
is the flexural rigidity per unit width in the joist
direction and D
s
the flexural rigidity per unit width in the slab
direction (including a correction for shear in open-web joists)
based on the moment of inertia of the uncracked concrete
(assume an average thickness t
c
for ribbed decks). The form
of Equation 7 is based on orthotropic plate action and the
factor 2 was determined by calibration to floor data as de-
scribed later. The effective panel width, B
j
, determined by
Equation 7 should be assumed to have an upper limit of
two-thirds of the total width of the floor perpendicular to the
joists or beams.
Where the beams or joists are continuous over their sup-
ports (including rolled sections shear connected to girder
webs), and an adjacent span is 0.7L
j
or greater, the effective
joist panel weight, W
j
, can be increased by 50 percent. The
reason for this increase is that continuity over supports en-
gages participation of adjacent floor panels in the fundamen-
tal mode of vibration. (Wyatt (1988) recommends an increase
of 70 percent where the adjacent span is 0.8L
j
or greater, 100
percent when it is 1.0L
j
.)
GIRDER PANEL MODE
The girder panel mode is associated with the natural fre-
quency of the girder alone. The natural frequency of this mode
can be estimated from
f
g
= 0.18 g /
g
(8)
Table 3.
Values of K and for use in Equation (4)
K
kN (kips)
Offices, residences, churches 58 (13.0) 0.03*
Shopping Malls 20 (4.5) 0.02
Footbridges 8 (1.8) 0.01
*0.05 for full-height partitions, 0.02 for floors with few non-structural com-
ponents (ceilings, ducts, partitions, etc.) as can occur in churches
Fig. 4. Proposed criterion for walking vibrations.
FOURTH QUARTER / 1993 121
where
g
is the deflection of individual girders relative to their
supports due to the weight supported. Composite action can
be assumed when the girders are directly connected to the
concrete slab, for example by welds to the steel deck. When
the girders are separated from the concrete slab by beams or
joist seats (shoes), they act as Vierendeel girders, i.e. partially
composite. It is recommended that the moment of inertia of
girders supporting joist seats be determined from:
I
g
= I
nc
+ (I
c
I
nc
) / 2 (9a)
for seat heights 75 mm (3 in.) or less, and
I
g
= I
nc
+ (I
c
I
nc
) / 4 (9b)
for seat heights 100 mm (4 in.) or more, where I
nc
and I
c
are
non-composite and fully composite moments of inertia re-
spectively. (These recommendations are subject to change
depending on the results of current research.) Normally the
girders are assumed to be simply supported unless dynamic
restraint is verified by analysis or experiment.
The mass weight of the girder panel mode can be estimated
from
W
g
= wB
g
L
g
(10)
where L
g
is the girder span and B
g
is the effective girder panel
width determined from
B
g
= 1.6 (D
j
/ D
g
)
1
4
L
g
(11)
where D
g
is the flexural rigidity per unit width in the girder
direction and D
j
the flexural rigidity per unit width in the joist
direction. Equation 11 is the same as Equation 7 except that
the factor 2 is reduced to 1.6 to take into account discontinuity
of joist systems over supports; if the joists consist of rolled
beams shear connected to girder webs the factor 1.6 can be
increased to 1.8. B
g
determined by Equation 11 should be
assumed to have a lower limit equal to the tributary panel
width supported by the girder and an upper limit of two-thirds
of the total floor width perpendicular to the girders.
Where the girders are continuous over their supports, and
an adjacent span is 0.7L
g
or greater, the mass weight, W
g
, can
be increased by 50 percent. This is due to participation of
adjacent floor panels, as discussed above for the joist panel
mode.
COMBINED MODE
Combined flexibilities of the joists and girders reduces the
natural frequency and makes the floor more susceptible to
noticeable walking vibration. For design purposes this can be
taken into account by a combined mode whose properties
may be estimated using the following interaction equations:
(i) The fundamental natural frequency can be approxi-
mated by the Dunkerly relationship:
f
o
= 0.18 g / (
j
+
g
) (12)
(ii) The equivalent mass weight can be approximated by
the interaction formula:
W =

j
+
g
W
j
+

j
+
g
W
g
(13)
If the girder span, L
g
, is less than the joist panel width, B
j
,
the combined mode is restricted and the system is effectively
stiffened. This can be accounted for by reducing the deflec-
tion,
g
, used in Equations 12 and 13 to

g
=
L
g
B
j

(
g
) (14)
where
0.5 L
g
/ B
j
1.0
EXAMPLE
Determine if the framing system for the typical interior bay
shown in Figure 5 satisfies the proposed criterion for walking
vibration. The structural system supports the office floors
without full-height partitions. For ease in reading, this exam-
ple will be carried out using Imperial units.
Concrete: 110 pcf, f
c
= 4,000 psi; n = E
s
/ 1.35 E
c
= 9.3
Deck thickness = 3.25 in. + 2 in. ribs = 5.25 in.
Deck weight = 42 psf
Beam Mode Properties
With an effective concrete slab width of 120 in. < 0.4 L
j
=
0.4 35 12 = 168 in., and considering only the concrete
above the steel form deck, the transformed moment of inertia
I
j
= 2,105 in.
4
For each beam
w
j
= 10(11 + 42 + 4 + 40 / 10) = 610 plf
which includes 11 psf live load and 4 psf for mechanical/ceil-
ing, and

j
=
5w
j
L
j
4
384EI
j
=
5 610 35
4
1,728
384 29 10
6
2,105
= 0.337 in.
The beam mode natural frequency from Equation 5 is:
f
j
= 0.18

386
0.337
= 6.09 Hz
Using an average concrete thickness, 4.25 in., the transformed
moment of inertia per unit width in the slab direction is
D
s
= 12 4.25
3
/ 12 9.3 = 8.25 in.
4
/ft
The transformed moment of inertia per unit width in the beam
direction is (beam spacing is 10 ft)
D
j
= 2,105 / 10 = 210.5 in.
4
/ft
The effective beam panel width from Equation 7 is:
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 122
B
j
= 2(8.25 / 210.5)
1
4
(35) = 31.3 ft
Since this is a typical interior bay, the actual floor width is at
least 3 30 = 90 ft, and
2

3
90 = 60 ft > 31.3 ft. Therefore,
the effective beam panel width is 31.3 ft.
The mass weight of the beam panel is from Equation 6,
adjusted by a factor of 1.5 to account for continuity:
W
j
= 1.5(610/10)(31.3 35) = 100,238 lbs = 100 kips
Girder Mode Properties
With an effective slab width of 0.4 30 12 = 144 in. and
considering the concrete in the form of deck ribs, the trans-
formed moment of inertia I
g
= 3,279 in.
4
For each girder
w
g
= 2.5 (610 35) / 30 + 50 = 1,829 plf,

g
=
5 1,829 30
4
1,728
384 29 10
6
3,279
= 0.350 in.
and
f
g
= 0.18

386
0.350
= 5.98 Hz
With D
j
= 210.5 in.
4
/ft and D
g
= 3,279 / 35 = 93.7 in.
4
/ft,
Equation 11 gives
B
g
= 1.8 (210.5 / 93.7)
1
4
(30) = 66.1 ft
which is less than
2

3
(3 35) = 70 ft. From Equation 10
W
g
= (1829 / 35)(66.1 30) = 103,626 lb = 104 kip
Combined Mode Properties
In this case the girder span (30 ft) is less than the beam panel
width (31.3 ft) and the girder deflection,
g
, is therefore
reduced according to 0.350 30 / 31.3 = 0.334 in. From
Equation 12,
f
o
= 0.18 386 / (0.337 + 0.334) = 4.32 Hz
and from Equation 13
W =
0.337
0.337 + 0.334
(100) +
0.334
0.337 + 0.334
(104) = 102 kips
For office occupancy without full-height partitions, = 0.03
from Table 3, thus
W = 0.03 102 = 3.06 kips
Evaluation
Application of Equations 4 for offices (see Table 3) results in
W = 3.06 kips > 13 exp (0.35 4.32) = 2.87 kips
or
f
o
= 4.32 Hz 2.86 ln (13 / 3.06) = 4.14 Hz
The floor is therefore judged satisfactory.
EDGE PANEL MODE
Unsupported edges of floors can cause a special problem
because of low-mass weight and sometimes decreased damp-
ing. Normally this is not a problem for exterior floor edges,
because of stiffening by exterior cladding or because walk-
ways are not located near exterior walls. Problems have
occurred, however, at interior floor edges adjacent to atria.
These edge members should often be made stiffer than current
practice suggests by use of the following assumptions in the
proposal criterion.
Where an interior edge is supported by a joist, the equiva-
lent mass weight of the joist panel can be estimated using
Equation 6 by replacing the coefficient 2 with 1 in Equation
7. Where an interior edge is supported by a girder, the equiva-
lent mass weight of the girder panel should be estimated on
the basis of the tributary weight supported by the girder. These
edge panels are then combined with their orthogonal panels
as recommended above.
CALIBRATION OF PROPOSED CRITERION
TO EXPERIENCE
The factor 2 in Equation (7) was determined by calibration to
data on one-way joist floor systems in Table 1 of Allen and
Rainer (1976). The results of applying the proposed criterion,
including recommended design parameters, to floors that
have been evaluated and tested is given in Tables 4 and 5.
Table 4 confirms application of the proposed criterion for
one-way systems, two-way systems, and interior edge panels.
Application of the CSA criterion (CSA, 1989) to the two-way
floor systems in Table 4, on the other hand, predicts that all
are satisfactory when in fact floors 12 and 13 are definitely
unsatisfactory. Table 5 confirms application of the proposed
criterion to two-way systems except for floor 3, a heavy floor
(3.6 kPa) with continuity in both directions. Two factors for
Fig. 5. Floor framing systemtypical interior bay.
FOURTH QUARTER / 1993 123
unsatisfactory performance of this floor are low damping
(criterion just met for = 0.015) and vibration transmission
due to girder continuity. Floors 7 and 10 are predicted to be
marginal.
The proposed criterion can also be compared to existing
criteria. Table 6 makes this comparison on the basis of mini-
mum values of W
j
for one-way beam or joist systems. The
basis for the values shown in Table 6 is given in Appendix III.
For office floors, Table 6 shows that all criteria are similar for
resonance with the third harmonic of the step frequency (5 to
7 Hz). This is not surprising because existing design criteria
are based to a large extent on experience with floors in the
frequency range 5 to 8 Hz.
The criteria, however, differ at other floor frequencies. The
CSA criterion is insufficient for frequencies less than 5 Hz
and conservative for frequencies beyond 7 Hz. The Murray
criterion has tendencies similar to the CSA criterion, but the
discrepancy with the proposed criterion is less severe. The
Wyatt criterion is close to the proposed criterion within a
broad frequency range, 3 to 8 Hz, but is more conservative
beyond 8 Hz.
For footbridges the proposed criterion is apparently a little
more conservative than the OHBDC (1983) criterion, but this
is offset by the difference in recommended values of (0.01
vs. 0.005 to 0.008 in the OHBDC). Third and fourth harmonic
resonance is not adequately considered by the OHBDC but
this is not serious in practice because footbridges with these
frequencies generally have sufficient mass to satisfy the pro-
posed criterion, Equation 4a.
Information on shopping centers is scarce. Application of
Equation 4a for shopping centers to the floor data in Cases 16
and 19 of Table 4, however, indicates agreement with user
reaction.
Tables 46, as well as Figure 3, therefore confirm the
applicability of the proposed criterion for walking vibration
to a wide variety of structures and occupancies.
Table 4.
Application of Proposed Design Criterion to Tested Floors
Case Reference or Location
Measured
Frequency
f
o
(Hz)
Span L (m)
Panel
Width B (m)
Equation 7
& 11
Damping
Ratio,
Table 3
W (kN)
User
Rating
2
Calc.
Criterion
Equation 4a
One-Way Joist Systems
1
2
3
4
5
6
7
8
9
10
11
Allen and Rainer (1976), #13
#9
#24
#5
#10
#2
#1
#18
#22
#19
#17
4.0
4.5
4.6
5.3
5.3
5.5
6.0
6.0
8.0
8.5
8.8
22.2
21.6
16.5
18.3
18.6
14.6
10.7
17.1
10.7
8.9
8.7
9.7
11.9
11.2
8.8
7.8
8.6
8.3
9.8
7.1
8.2
7.6
0.03
0.03
0.03
0.015
0.015
0.03
0.03
0.015
0.03
0.015
0.015
19.3
26.9
16.6
6.0
5.4
9.4
6.6
7.5
5.5
3.3
2.5
14.3
12.0
11.6
9.1
9.1
8.5
7.1
7.1
3.5
3.0
2.7
S
S
S
U
U
S
U
B
S
B
U
Two-Way JoistGirder Systems
12
13
14
15
16
Quebec City
Quebec City
Quebec City
Matthews, et al (1982)
Pernica and Allen (1982)
4.5
5.4
7.2
6.2
5.2
(7.6, 7.6)
3
(7.6, 7.6)
(7.6, 7.6)
(9, 12.5)
(7.6, 12.2)
(9.1, 11.9)
3
(9.1, 8.6)
(7.4, 10.7)
(9.7, 11.3)
(8.1, 15.0)
0.03
0.03
0.03
0.03
0.02
6.2
5.4
5.2
9.5
11.8
12.0
8.8
4.7
6.6
3.2
Very U
U
S
S
S
Interior Edge Panels
17
18
19
Quebec City
Edmonton
Pernica and Allen (1982)
5.1
5.1
5.6
13.7
17.5
12.2
2.3
4.6
3.3
0.03
0.03
0.02
2.4
8.4
2.5
9.7
9.7
2.8
Very U
U
U
Notes:
1
K = 58 for all cases except #16 and #19, where K = 20 applies
2
U = unsatisfactory, S = satisfactory, B = borderline
3
The first entry inside the brackets refers to the joist panel, the second refers to the girder panel
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 124
NATURAL FREQUENCIES GREATER THAN 9 HZ
When the natural frequency is greater than 9 Hz, harmonic
resonance does not occur, but walking vibration can still be a
problem. Because the natural frequencies are high compared
to the main loading frequencies, the floor response is gov-
erned primarily by stiffness relative to a concentrated load.
Experience indicates a minimum stiffness of approximately
1 kN per mm (5.7 kips per in.) deflection for office and
residential occupancies.
For light floors with natural frequencies in the range 9 to
18 Hz there may also be adverse reaction to floor motion
caused by step-impulse forces. Experience indicates that ad-
verse reaction to step impulses depends primarily on mass
(initial floor velocity equals impulse divided by mass) and
vibration decay time, the shorter the decay time the better.
The decay time decreases in proportion to clamping ratio
times floor frequency. Wyatt (1989) recommends an impulse
criterion beyond 7 Hz floor frequency, but beyond approxi-
mately 9 Hz the criterion becomes overly conservative be-
cause it ignores the benefits of decreased decay time. Ohlsson
(1988) recommends an impulse criterion which takes decay
time into account, but the criterion is complex for design. The
resonance criterion, Equation 4a, is in a form that correctly
reflects impulse discomfort except that the right-hand side has
not been correctly determined. If, however, Equation 4a with
K = 58 for office floors is extended beyond 9 Hz, it decreases
rapidly until approximately 18 Hz when the stiffness criterion
of 1 kN/mm (5.7 k/in.) starts to control the design of the floor.
Application of Equation 4a to the examples in Ohlsson (1988)
also indicates that it gives a reasonable evaluation for floors
between 9 and 18 Hz.
To ensure satisfactory performance of office and residen-
tial floors with frequencies greater than 9 Hz it is recom-
mended that Equations 4 be used in conjunction with the
stiffness criterion of 1 kN/mm (5.7 k/in.).
CONCLUSION
Walking forces produce motions which are related to reso-
nance, impulse response, and static stiffness. Resonance con-
trols the design of floors and footbridges with natural frequen-
cies less than approximately 9 Hz, static stiffness controls the
design of floors with frequencies greater than approximately
18 Hz, and impulse response controls the design of floors with
frequencies in between.
A simple criterion for resonance vibration of floor and
footbridge structures, Equations 4, is proposed for design,
along with a recommended procedure for determining the
Table 5.
Application of Proposed Design Criterion to Floors Investigated by Murray (1981)
1
Case
Calculated
Frequency f
o
(Hz)
Span (m) Panel Width B(m)
Estimated
Damping
Ratio,
Table 3
W (kN)
User
Rating
Beam Girder
Joist or
Beam Girder Calculated
Criterion
Equation
(4a)
1 7.0 10.5 6.0 6.3 11.1 0.015 2.0 4.9 U
2 7.0 10.5 6.0 6.3 11.1 0.05 6.6 4.9 S
3 4.0 7.3 12.2
2
9.2
2 2
16.9
2
0.02 18.6 14.1 U
4 7.7 7.0 7.2 7.2 9.0 0.015 2.0 3.9 U
5
3
5.9 12.2 Wall 7.3 ? 0.015 3.5 7.3 U
6 5.9 12.2 Wall 7.3 ? 0.05 11.7 9.2 S
7 5.3 13.4 6.4 8.0 19.8 0.03 9.3 9.2 U
8 6.1 9.1 6.1 6.4 11.6 0.03 3.9 6.9 U
9 5.1 5.5 12.5 6.5 9.1 0.03 6.7 9.6 U
10 5.2 11.6 9.8
2
8.7
2
19.9 0.02 9.7 9.4 U
11
3
6.4 12.2 ? 7.9 ? 0.02 5.1 6.1 U
Notes:
1
All open web joist on girder systems except #3 and #10 (beams shear connected to girders)
2
Members continuous over supports (Wj or Wg increased by 1.5)
3
Joist systems supported on stiff girders, frequency fo estimated from fj
FOURTH QUARTER / 1993 125
required floor properties. The proposed criterion, based on
acceptable vibration for human reaction, compares well with
existing criteria and is confirmed by experience with tested
floors. Recommended values of the criterion parameters,
however, are expected to be improved by further experience
and research.
Floors of offices and residential occupancies with frequen-
cies greater than 9 Hz should also be checked both for a
minimum static stress under concentrated load of 1 kN/mm
(5.7 kips/in.) and for impulse response by means of Equa-
tions 4.
APPENDIX I: REFERENCES
1. American Institute of Steel Construction, Specification
for Structural Steel BuildingsAllowable Stress Design
and Plastic Design, AISC, Chicago, 1989.
2. Allen, D. E. and Rainer, J. H., Vibration Criteria for
Long-Span Floors, Canadian Journal of Civil Engineer-
ing, 3(2), June, 1976, pp. 165171.
3. Bachmann H., Case Studies of Structures with
Man-Induced Vibrations, Journal of Structural Engi-
neering, ASCE, Vol. 118, No. 3, 1992, 631647.
4. British Standard BS5400, Part 2: Steel, Concrete and
Composite Bridges: Specification for Loads, Appendix C,
British Standards Institution, 1978.
5. Canadian Standard CAN3-S16. 1-M89: Steel Structures
for BuildingsLimit States Design, Appendix G: Guide
for Floor Vibrations, Canadian Standards Association,
Rexdale, Ontario, 1989.
6. International Standard ISO 2631-2, Evaluation of Human
Exposure to Whole-Body VibrationPart 2: Human Ex-
posure to Continuous and Shock-Induced Vibrations in
Buildings (1 to 80 Hz), International Standards Organiza-
tion, 1989.
7. International Standards ISO 10137, Basis for the Design
of StructuresServiceability of Buildings Against Vibra-
tion, International Standards Organization, 1992.
8. Lenzen, K. H., Vibration of Steel Joists, Engineering
Journal 3(3), 1966, pp. 133136.
9. Matthews, C. M., Montgomery, C. J., and Murray, D. W.,
Designing Floor Systems for Dynamic Response,
Structural Engineering Report No. 106, Department of
Civil Engineering, University of Alberta, Edmonton, Al-
berta, 1982.
10. Murray, T. M., Acceptability Criterion for Occupant-In-
duced Floor Vibrations, Engineering Journal, 18(2),
1981, 6270.
11. Murray, T. M., Building Floor Vibrations, Engineering
Journal, Third Quarter, 1991, 102109.
12. Ontario Highway Bridge Design Code, Ontario Ministry
of Transportation and Communication, Toronto, 1983.
13. Ohlsson, S. V., Ten Years of Floor Vibration Research
A Review of Aspects and Some Results, Proceedings of
the Symposium/Workshop on Serviceability of Buildings.
Vol. I, Ottawa, 1988, pp. 435450.
14. Pernica, G., and Allen, D. E., Floor Vibration Measure-
ments in a Shopping Centre, Canadian Journal of Civil
Engineering, 9(2), 1982, pp. 149155.
15. Rainer, J. H., Pernica, G., and Allen, D. E., Dynamic
Loading and Response of Footbridges, Canadian Jour-
nal of Civil Engineering, 15(1), 1988, pp. 6671.
16. Tredgold, T., Elementary Principles of Carpentry, 2nd
Ed., Publisher unknown, 1828.
17. Wyatt, T. A., Design Guide on the Vibration of Floors,
Steel Construction Institute Publication 076, London,
1989.
APPENDIX II: NOTATION
The following symbols are used in this paper:
a = acceleration
a
o
= acceleration limit
B = effective width of a panel
D = flexural rigidity or transformed moment of inertia
per unit width of a panel
Table 6.
Comparison of Various Design Criteria for Walking Vibrations
Natural
Frequency
f
o
(Hz)
Minimum Value of Damping Ratio Times Effective Mass Weight, W
j
(kN)
Offices, Residences Footbridges
Equation 4a CSA (1989)
1
Murray (1981)
1
Wyatt (1989) Equation 4a OHBDC (1983)
2
4
6
8
10
28.8
14.3
7.1
3.5
1.75
NA
4
6
7
8
NA
5.87.6
5.87.6
5.87.6
5.87.6
NA
17.5
8.8
1
3.0
1
1
3.0
1
4.0
2.0
1.0
0.5
0.24
3
1.8

Note:
1
Results are given for a standard case of finished floor without full-height partitions ( = 0.03)
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 126
E
s
= modulus of elasticity for steel
f = step frequency
f
j
= natural frequency of joist or beam panel
f
g
= natural frequency of girder panel
f
o
= fundamental natural frequency of floor structure
g = acceleration due to gravity; subscript indicating
girder
i = ith harmonic of step frequency
j = subscript indicating joint or beam
K = factor in Equation 4 taking into account occupant
sensitivity to vibration
L = span of joint, beam, or girder (with subscript j
or g)
P = weight of a person (0.7 kN assumed)
R = reduction factor in Equation 2
W = effective mass weight of floor vibrating in the
fundamental mode
w = unit weight of floor panel, including acting live
load
w
j
or w
g
= unit weight of joist or girder per unit length

i
= dynamic load factor for ith harmonic of step fre-
quency
= damping ratio
= deflection of member under weight supported
APPENDIX III: BASIS FOR COMPARISON OF
VIBRATION CRITERIA
Existing design criteria for walking vibration can be com-
pared with the proposed criterion by considering a standard
joist or beam panel on stiff supports. To make a valid com-
parison, each criterion must be considered as a total package.
This requires adjustments to the criteria to take account of
differences in the form of the design equations and in the
recommended values of design parameters. To make a com-
parison, all criteria will be transformed to a common measure
W
j
as defined for the proposed criterion.
The following frequency relationship for a simply sup-
ported joist panel will be used to transform all criteria to the
common measure, W
j
:
f
o
=

gD
j

wL
j
4
(A1)
where w is the unit weight of the panel
Canadian Standards Association (CSA, 1989)
This criterion has been used in Canada since 1975, with minor
modifications in 1984. For the standard joist panel, the CSA
criterion can be expressed as follows:
w(40t
c
)L
j
(kN) > 0.6f
o
/ (a
o
/ g) (A2)
where t
c
is the effective concrete thickness, 40t
c
is the effec-
tive slab width, and a
o
/ g is a limiting heel-impact accelera-
tion determined from Figure 6. Equation A2 can be expressed
in terms of W
j
if a correction is made for the effective panel
width. For a typical case of a 5.5 Hz floor, span L
j
= 12 m and
concrete thickness t
c
= 75 mm, application of Equation 7
results in an effective width of 8.3 m or 110t
c
compared to
40t
c
in Equation A2. If Equation A2 is multiplied by
110 / 40 it becomes
W
j
(kN) > 1.65f
o
/ (a
o
/ g) (A3)
Minimum values of W
j
for the CSA criterion in Table 6
were determined from Equation A3 using the criterion for
finished floors in Figure 6 and = 0.03 from Table 3.
Murray (1981)
On the basis of a review of field data from 91 floors, Murray
(1981) recommended the following criterion, presently
widely used in the U.S.:
> 0.35A
o
f
o
+ 0.025 (A4)
where A
o
is the initial amplitude of vibration (inches) due to
Fig. 6. Annoyance criteria for floor vibrations in residential,
school, and office occupanices (CSA, 1989).
FOURTH QUARTER / 1993 127
a standard heel impact. Equation A4 is plotted in Figure 7
along with the floor data. To determine A
o
Murray provides
the following expression for a simply supported one-way
floor system:
A
o
= DLF
s
(A5)
where
s
is the static deflection of the joist panel under a
concentrated load of 600 lb. and DLF is a dynamic load factor
to obtain the maximum amplitude of vibration for a standard
heel impact. DLF ranges from 0.15f
o
at f
o
= 4 Hz to 0.12f
o
at
f
o
= 10 Hz, and can therefore be approximated by 0.14f
o
, its
value at f
o
= 6 Hz. Thus,
A
o
=
(1.14f
o
) 600L
j
3
48D
j
B
M
(A6)
where B
M
is the effective joist panel width as defined later.
Substitution of Equation A6 in Equation A4 after elimination
of D
j
by means of Equation A1 results in the following
criterion:
>
584
wB
M
L
j
+ 0.025 (A7)
For the standard case of finished office floor without full-
height partitions, = 0.03 according to Table 3 and = 0.045
according to Murray. For this case Equation A7 becomes
wB
M
L
j
= 584 / (0.045 0.025) = 29,200 (A8)
Murray (1991) provides expressions for determining B
M
in
terms of beam or joist spacing times the number of effective
joists. Two expressions are used, one for normal hot-rolled
beam (spacing more than 30 in.); the other for closely spaced
joists (30 in. or less). The expression for narrow spacing is
equivalent to
B
M
=
32

D
S
D
j

1
4
L
j
= 1.35

D
S
D
j

1
4
L
j
= 0.675B
j
(A9)
where B
j
is defined according to Equation 7, and the expres-
sion for wide spacing can be approximated by
B
m
= 1.03

D
S
D
j

1
4
L
j
= 0.515 B
j
(A10)
Substitution of Equations A9 or A10 in Equation A8 results
in minimum values of W
j
equal to 43,260 lb for narrow
spacing and 56,700 lb for wide spacing. For the standard case,
= 0.03, the corresponding minimum values of W
j
included
in Table 6 are 1,300 lb (5.8 kN) and 1,700 lb (7.6 kN).
Wyatt (1989)
Wyatt (1989) proposed two design criteria for office floors,
one a resonance criterion for floor frequencies up to 7 Hz, the
other an impulse response criterion for floor frequencies
greater than 7 Hz. For the one-way beam or joist system, the
resonance criterion can be expressed (with rearrangement and
change of symbols) as
(wB
w
L
j
) > 667C
f
/ F (A11)
where C
f
is a loading coefficient (0.4 for second harmonic
loading and 0.2 for third harmonic loading), F is a rating
factor which depends on the office environment (12 for a busy
office, 8 for a general office, and 4 for a special office) and
B
w
is the joist panel width. For the one-way system Wyatt
recommends
B
W
= 4.5

gD
S
f
o
2
w

1
4
(A12)
which can be expressed in the same form as Equation 7 by
use of Equation A1. After substitution, B
w
in Equation A12
becomes equal to 1.8B
j
, where B
j
is defined by Equation 7.
Wyatt, however, recommends a concrete modulus elasticity
25 percent higher than recommended for D
s
in Equation 7.
With this correction B
w
, becomes equal to 1.9B
j
. Equation A11
can therefore be expressed as
W
j
> 351C
f
/ F (A13)
for floor frequencies below 7 Hz. Table 6 contains minimum
values of W
j
assuming F = 8 for a general office.
For floor frequencies greater than 7 Hz Wyatt recommends
the following impulse criterion:
wSL
j
> 294 / F (A14)
where S is the member spacing. Equation A14 may be ex-
pressed in terms of W
j
if it is multiplied by B
j
/ S. Based on
Fig. 7. Murray criterion, Equation (A4), compared
to floor data (Murray 1981).
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 128
an assumed beam spacing of 2.5 m used in Wyatts examples
and a typical value B
j
= 6.8 m for an 8 Hz floor of span 9 m
and concrete thickness of 75 mm. Equation A14 can be
approximated by
W
j
> 800 / F (A15)
Table 6 contains a minimum value of Equation A15 at
f
o
= 8 Hz for a general office floor for which F = 8 and =
0.03.
FootbridgesOntario Highway Bridge Design Code
(OHBDC, 1983)
The OHBDC (1983) design criterion for footbridges is based
on a pedestrian or jogger exerting a dynamic force of
P cos 2ft where P is 0.7 kN, = 0.257 and f, the step
frequency, takes on any value between 1 and 4 Hz. The
footbridge is modeled as an SDOF beam which vibrates at the
first flexural frequency, f
o
. For a simply supported footbridge,
the resonance response for flexural frequencies up to 4 Hz can
be determined from Equation 2 with a value of R which is
determined by the length of the footbridge. If, for a typical
case R is assumed equal to 0.7, the maximum acceleration is
determined from
a
max
/ g = 0.7(0.257)0.7 / W
j
= 0.126 W
j
(A17)
where W
j
is the weight of the footbridge. The OHBDC rec-
ommends limiting values of a
max
/ g equal to 0.042 at f
o
= 2 Hz
and 0.072 at f
o
= 4 Hz. Thus Equation A17 can be inverted to
a criterion for minimum value of W
j
equal to 0.126 / 0.042
= 3 kN at f
o
= 2 Hz and 0.126 / 0.072 = 1.8 kN at f
o
= 4 Hz.
For a flexural frequency beyond 4 Hz, the OHBDC gives
an incorrect assessment because it neglects resonance with
the higher harmonics of the walking and jogging forces.
FOURTH QUARTER / 1993 129
ABSTRACT
A practical LRFD-based analysis method for the design of
semi-rigid frames is proposed. The proposed method uses
first-order elastic analysis with a notional lateral load for the
second-order effects. In the proposed method, a simplified
three-parameter model describing the tangent rotational stiff-
ness of semi-rigid connections is used.
1. INTRODUCTION
Although partially restrained (PR) construction is permitted
by the AISC Specification for Structural Steel Buildings
Load and Resistance Factor Design, no specific analysis or
design guidance is given in the current LRFD and ASD
specifications for these partially restrained frames.
Recently, a simplified procedure for the analysis and de-
sign of semi-rigid frames was proposed by Barakat and
Chen,
1
using the B
1
and B
2
amplification factors together with
the beam-line concept. However, the beam-line method can
not adequately predict the drift of unbraced frames and the
calculation of effective length factor is cumbersome and
time-consuming.
A simplified procedure to improve these drawbacks is
introduced in this paper. Here, as in the Barakat method, the
proposed method is based on first-order linear elastic analy-
sis, but the second-order effect will be included with the use
of notional lateral loads.
2. MODELING OF SEMI-RIGID CONNECTIONS
2.1 Connection Models
Most existing connection models express the moment in
terms of rotation from which the tangent stiffness can be
derived. This paper proposes a direct tangent-stiffness expres-
sion for flexible connections. This proposed tangent-stiffness
model is based on the concept that connection stiffness de-
grades gradually from an initial stiffness, K
i
, to zero following
a nonlinear relationship of the simple form:
dM
d
r
K
t
K
i

M
M
u
_

,
c
1
1
]
(1)
where
K
t
= tangent stiffness
K
i
= initial connection stiffness
M
u
= ultimate bending moment capacity
M = connection moment
C = shape factor account for decay rate of K
t
, C > 0
The moment-rotation (M
r
) behavior of bolted extended
end-plate beam-to-column connections tested by Yee and
Melchers
2
is compared with the proposed model in Figure 1
and a good agreement is observed with C = 1.6. In Figure 1,
the initial stiffness, K
i
= 546,666 in-kip/rad, is the tangent to
the starting point of the curve. The ultimate moment capacity,
M
u
= 3,539 in-kip, is determined by test. The value C is used
to control the shape of a convex curve. If C is equal to 1, K
t
decreases linearly. When C is less than 1, K
t
decreases more
rapidly. If C is greater than 1, K
t
decreases much slower. This
is illustrated in Figure 1 with C = 1.0, 1.6, and 2.2 respectively.
In the following, the proposed tangent stiffness connection
model will be applied to several types of connections, includ-
ing the extended end-plate, top and seat angle with double
web angles, framing angles, and single-plate connections.
LRFD Analysis for Semi-Rigid Frame Design
WON-SUN KING AND WAI-FAH CHEN
Won-Sun King is associate professor, Department of Civil
Engineering, Chung Cheng Institute of Technology, Ta-Hsi,
Tao-Yuan, Taiwan.
Wai-Fah Chen is professor and head of Structural Engineer-
ing, School of Civil Engineering, Purdue University.
Fig. 1. Moment-rotation curves of Yee connection (1986).
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 130
The connections ranged from very stiff to rather soft connec-
tions. The moment-rotation curve is obtained by numerical
integration of tangent-stiffness Equation 1.
(a) Jenkins Bolted Extended End Plate
A comparison of the proposed model with one of the Jenkins,
Tong, and Prescott extended end-plate connection test
3
is
shown in Figure 2. In this test, the beam is 305165 UB54,
the column is 254254 UC132 (stiffened), bolts are M20
grade 8.8, thickness of end plate is 20 mm and K
i
= 786,732
in.-kip/rad and M
u
= 1,989 in.-kip. Good agreement is ob-
served with C = 0.555.
(b) Bolted Top and Bottom Angles with Web Angles
Azizinamini, Bradburn, and Radziminski
4
reported test re-
sults on bolted semi-rigid steel beam-to-column connections.
These connections are comprised of top and bottom angles
connected to the flanges along with web angles. ASTM A36
steel was used for the members and the connection elements.
Eighteen specimens were tested. The beam tested was a
W1438, the bolt diameter is 22.2 mm, and the web angles
are 2L43
1

4
. The thickness of flange angles is 15.9 mm,
and the length of the test beam is 203.2 mm. The test number
14S8 with K
i
= 677,025 in.-kip/rad and M
u
= 1,707 in.-kip
compares well with that of the proposed model in Figure 3
with C = 0.34.
(c) Bolted Framing Angles
Bolted double-web angles were tested by Lewitt, Chesson,
and Munse at the University of Illinois. In 1987, Richard, et
al
5
proposed a four-parameter formula to describe these full-
scale tests. Figure 4 compares the results of the proposed
model with one of these tests using a five-bolt design with
rivets in the angle-to-beam web connection with K
i
= 206,667
in.-kip/rad and M
u
= 761 in.-kip.
(d) Single Plate
A total of seven tests were made by Richard
6
on single-plate
connections. The first set of two-, three-, five-, and seven-bolt
tests were run with the framing connection plate welded to a
flange plate which was in turn bolted to the support column.
A second set of tests was run on the two-, three-, and five-bolt
connections with the framing connection plate welded to the
support column. In these tests, three bolts were used to
connect the beam and the single plate. The bolts are A325
3

4
-in. diameter, and the plate thickness is
3

8
-in. The moment-
rotation curve of the proposed model compares well with one
of the tests as illustrated in Figure 5 with K
i
= 51,000 in.-
kip/rad, M
u
= 137 in.-kip, and C = 0.22.
2.2 Initial Stiffness
For simplicity, researchers
7,8,9
have been using the initial
connection stiffness, K
i
, for their semi-rigid frames analysis.
The use of initial stiffness throughout the flexible frame
analysis results in a frame behavior that is generally too stiff
when the frame is subjected to a normal loading condition.
Extensive studies of frames by Ackroyd
10
with nonlinear
connections indicate that the secant stiffness of beam-to-col-
umn connections near ultimate frame capacity was typically
20 percent of the initial stiffness, K
i
, at leeward ends of girders
and 80 percent of K
i
at the windward ends of girders, when
the frame is subjected to combined gravity and wind loading.
It seems, therefore, reasonable to use an average connection
stiffness of 0.5K
i
when computing the design moments. This
is adopted in the present analysis.
3. DESIGN FORMULA IN AISC-LRFD
The equation for the maximum strength of beam-columns is
given by AISC-LRFD as
Fig. 2. Moment-rotation curves of Jenkins connection (1986).
Fig. 3. Moment-rotation curves of Azizinamini
connection (1987).
FOURTH QUARTER / 1993 131
for
P
u

c
P
n

0.2,
P
u

c
P
n

+
8
9

M
ux

b
M
nx
+
M
uy

b
M
ny
_

,
1.0 (2)
for
P
u

c
P
n

< 0.2,
P
u

2
c
P
n

+

M
ux

b
M
nx
+
M
uy

b
M
ny
_

,
1.0 (3)
where
P
n
= ultimate compression capacity of an axially
loaded column
M
nx
, M
ny
= ultimate moment-resisting capacity of a
laterally unsupported beam about x and y axes,
respectively

c
= column resistance factor (= 0.85)

b
= beam resistance factor (= 0.9)
P
u
= design axial force
M
ux
, M
uy
= member design moment about x and y axes,
respectively, with:
M
u
B
1
M
nt
+ B
2
M
lt
(4)
M
nt
= first-order moment in the member assuming no
lateral translation in the frame
M
lt
= first-order moment in the member as a result of
lateral translation of the frame
B
1
= P- moment amplification factor
B
1

C
m
1
P
u

P
e

1 (5)
B
2
= P- moment amplification factor
B
2

1
1
P
u

o
HL
(6)
C
m
= 0.6 0.4M
1
/ M
2
, where M
1
/ M
2
is the ratio of the
smaller to the larger end moment of a member
P
e
=
2
EI / (KL)
2
P
u
= axial loads on all columns in a story

o
= first-order translational deflection of the story under
consideration
H = sum of all story horizontal forces producing
o
L = story height
K = effective length factor determined from the
alignment chart
The second-order effects are taken into account approxi-
mately by the moment amplification factors B
1
and B
2
on the
nonsway and sway moments obtained from first-order elastic
analyses, respectively. It usually leads to conservative results.
4. BEAM-COLUMN STIFFNESS IN
SECOND-ORDER ELASTIC ANALYSIS
For second-order elastic analysis, we use the usual element
geometric stiffness matrix combined with the update of the
element geometry during the analysis. The first three terms in
the Taylor series expansion of the elastic stability functions
are retained for the axial compressive force P to increase the
accuracy of the element stiffness. The corresponding terms in
stiffness matrix were obtained by Goto and Chen
11
as
K
ii

4EI
L
+
2PL
15
+
44P
2
L
3
25,000EI
(7)
K
ij

2EI
L

PL
30

26P
2
L
3
25,000EI
(8)
Fig. 4. Moment-rotation curves of Lewitt
connection (Richard, 1987). Fig. 5. Moment-rotation curves of Richard connection (1980).
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 132
K
ji

2EI
L

PL
30

26P
2
L
3
25,000EI
(9)
K
jj

4EI
L
+
2PL
15
+
44P
2
L
3
25,000EI
(10)
where
P = axial force in member
A = area of a cross section
r
i
= internal reactions at both ends of a member = 1,6
d
i
= displacements at both ends of a member = 1,6
The beam stiffness matrix in Equation 11 can be simplified
by recognizing that the axial force in beams of rectangular
frames is usually negligible. That is, by setting K
ii
= K
jj
=
4EI / L, K
ij
= K
ji
= 2EI / L, and P = 0.
The stiffness matrix of a beam-column can be modified to
include the effect of semi-rigid connections by combining the
member stiffness with the connection stiffness using a static
condensation. Details of this procedure are given in Chen and
Lui,
12
and the resulting member stiffness matrix has the form:

'

;
)


r
1
r
2
r
3
r
4
r
5
r
6
AE
L
0 0
AE
L
0 0

'

;
)

d
1
d
2
d
3
d
4
d
5
d
6
(11)
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2
+
P
L
(K
ii
+ K
ji
)
L
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2

P
L
(K
ij
+ K
jj
)
L
0
(K
ii
+ K
ij
)
L
K
ii 0
(K
ii
+ K
ij
)
L
K
ij
AE
L
0 0
AE
L
0 0
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2

P
L
(K
ii
+ K
ji
)
L
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2
+
P
L
(K
ij
+ K
jj
)
L
0
(K
ji
+ K
jj
)
L
K
ji 0
(K
ji
+ K
jj
)
L
K
jj

'

;
)


r
1
r
2
r
3
r
4
r
5
r
6
AE
L
0 0
AE
L
0 0

'

;
)

d
1
d
2
d
3
d
4
d
5
d
6
(12)
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2
+
P
L
(K
ii
+ K
ji
)
L
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2

P
L
(K
ij
+ K
jj
)
L
0
(K
ii
+ K
ij
)
L
K
ii
0
(K
ii
+ K
ij
)
L
K
ij

AE
L
0 0
AE
L
0 0
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2

P
L
(K
ii
+ K
ji
)
L
0
(K
ii
+ 2K
ij
+ K
jj
)
L
2
+
P
L
(K
ij
+ K
jj
)
L
0
(K
ji
+ K
jj
)
L
K
ji
0
(K
ji
+ K
jj
)
L
K
jj

FOURTH QUARTER / 1993 133


where
K
ii

S
ii
+
S
ii
S
jj

R
j

S
ij
S
ij

R
j
1
1
]

1
R
(13)
K
jj

S
jj
+
S
ii
S
jj

R
i

S
ij
S
ij

R
i
1
1
]

1
R
(14)
K
ij
K
ji

S
ij

R
(15)
The coefficients R
i
and R
j
in Equations 13 and 14 are the
instantaneous tangent stiffness coefficients of the connections
at ends i and j of the member respectively. These coefficients
are obtained from Equation 1 when the connection is in the
state of loading, and are set equal to K
i
when the connection
is in the state of unloading. Also, the parameter R* is given
by
R

1 +
S
ii

R
i
_

1 +
S
jj

R
j
_

,

S
ij
S
ij

R
i
R
j
(16)
S
ii
S
jj
K
ii
K
jj
(17)
S
ij
S
ji
K
ij
K
ji
(18)
in which P is negative for compressive force, and is small or
zero for beam elements and can be neglected.
5. THE PROPOSED METHOD
Several simplifications are made in the present formulation.
The moments of beam-column joints must be less than the
ultimate moment M
u
of semi-rigid connections or the plastic
moment capacity M
pc
of beam-columns. The combined axial
load and end moments in any member must satisfy the AISC-
LRFD bilinear interaction equations.
5.1 Rigid Frame Analysis
1. Perform the first-order elastic rigid frame analysis.
2. Compute notional lateral loads, H, using the relation-
ship
S
F

H

o

H + P
u
/ L

(19)
where
o
is the first-order translational deflection of the story,
and is the second-order translational deflection of the story
under consideration. From Equation 19, we have
(H + P
u
/ L)

o
H

1 +
P
u

HL
_

,

o
(20)
from which we obtain

1
P
u

o
HL
_

,

o
(21)
or

1
P
u

o
HL
_

,
B
2

o
(22)
The is the second-order lateral deflection due to P-
effect, and the notional lateral load is defined as
H H + P
u
/ L (23)
3. Use H and original gravity loads to perform first-order
elastic rigid frame analysis. The results of this step
include the second-order effect.
4. Calculate B
1
factor with the effective length factor K =
1.0 for each column and multiply the corresponding end
moments.
5. Check the AISC-LRFD bilinear interaction equations.
5.2 Semi-Rigid Frame Analysis
1. Select connections from the maximum beam-column
joint moments in rigid frame analysis.
2. Determine the initial stiffness, K
i
, of connections from
test results or any other available methods.
3. Substitute 0.5K
i
of connection stiffness for the semi-
rigid joint. The average connection stiffness 0.5K
i
as
suggested by Ackroyd
10
is adopted here.
4. Use 0.5K
i
for semi-rigid connection stiffness with the
notional lateral loads H to carry out the first-order
elastic analysis.
5. Calculate B
1
factor with effective length factor K = 1.0
for P
e
of each column and multiply the corresponding
larger end moments.
6. Check the AISC-LRFD bilinear interaction equations.
The effective length factor, K, for the column strength,
P
n
, has to be modified in the case of semi-rigid frames.
For beams connected to columns with semi-rigid con-
nections rotational stiffness, K
i
, at both ends, a simple
modification of the relative stiffness, G, factors with the
modified moment inertia of beam is (Chen and Lui
13
):
I
I
1 +
2EI
K
i
L
(24)
The I is used in G factors for the determination of the
effective length factor, K, for the value of F
cr
which is the
critical column stress.
6. NUMERICAL EXAMPLES
The proposed method will now be illustrated by numerical
examples. Comparisons are made between results using di-
rect second-order elastic analysis, Barakats method,
1
and the
proposed method. The semi-rigid frame examples include
single-story and multi-story frames. All examples are ana-
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 134
lyzed with a personal computer. All beams subjected to uni-
formly distributed loads are divided into two equal elements.
6.1 Two-Story One-Bay Frame with Concentrated
Loads
The two-story one-bay frame as shown in Figure 6 is analyzed
with both rigid and semi-rigid connections. Two lateral loads,
H, and four constant concentrated gravity loads, P, of 100 kips
are applied at the beam-column joints of the frame. The
flexible connections used are shown in Figure 2 where M
u
is
less than the plastic moment M
p
of beams and columns. The
0.5K
i
of Jenkins connection is 393,366 in-kip/rad. The sec-
ond-order lateral displacement at Joint 5 is

5
=

o5

1
P
u

o5
HL

=
1.512

1
2 100 1.512
2 10 144

= 1.69 in.
The second-order lateral load at Joint 5 is
H
5
= H
5
+ P
u

5
/L = 10 + 2 100 1.69/144 = 12.35 kips
The second-order lateral displacement at Joint 3 is

3
=

o3

1
P
u

o3
HL

=
1.01

1
4 100 1.01
2 10 144

= 1.17 in.
The notional lateral load at Joint 3 is
H
3
= H
3
+ P
u

3
/L = 10 + 4 100 1.17/144 = 13.25 kips
All moments of beams and columns predicted by the
proposed method are normalized with respect to that of
second-order elastic analysis and are summarized in Tables 1
and 2. The lateral displacements at windward beam-column
joints are shown in Table 3. The mean values are the sum of
normalized values of each member divided by the number of
total members. All the results predicted by the proposed
method are close to the exact solutions. It is found that the
maximum moment and lateral displacement can be predicted
well by the proposed method.
6.2 Two-Story One-Bay Frame with Uniformly
Distributed Loads
A two-story one-bay frame used by Barakat,
1
et al as shown
in Figure 7(a) is employed here for comparison of the maxi-
mum moments in members. The semi-rigid connection la-
beled III-17 is shown in Figure 8 and compared with the
proposed connection model. The moments predicted by sec-
ond-order elastic analysis the Barakat method, and the pro-
posed method are compared in Table 4. The average value of
Column 3 in Table 4 is 0.98, while the average value of
Column 5 is 0.97. The Barakat method is slightly less conser-
vative in this example.
To verify the validity of the proposed method for soft
semi-rigid connections, the bolted framing angles tested by
Lewitt
5
are used. The lateral loads and uniformly distributed Fig. 6. Two-story one-bay frame with concentrated loads.
Table 1.
Maximum Moments in Elastic Rigid
Frame Analysis (in.-kips)
(Two-story one-bay frame, Figure 6)
Element
No.
(1) (2) (3) (4)
First-Order
(Exact)
Second-Order
(Exact) Proposed (3) / (2)
1 1449 1649 1839 1.12
2 712 794 894 1.13
3 1443 1670 1847 1.11
4 1437 1664 1839 1.11
5 711 794 893 1.13
6 712 794 894 1.13
FOURTH QUARTER / 1993 135
loads are reduced as shown in Figure 7(b), so the maximum
moments in members of the two-story one-bay frame are less
than the ultimate moment, M
u
, semi-rigid connections. The
results predicted by the proposed method are compared with
that of second-order elastic analysis in Table 4(b) and the
lateral displacements are shown in Table 4(c). It can be
concluded that the proposed method is valid for soft connec-
tions, although the second-order lateral loads in the proposed
method are determined from a rigid frame.
6.3 Three-Story One-Bay Frame with Uniformly
Distributed Loads
The three-story one-bay frame shown in Figure 9 is analyzed
with semi-rigid connections labeled III-17. Three beam-col-
umn joints are subjected to concentrated lateral loads. All the
beams are subjected to uniformly distributed gravity loads.
The results by the Barakat method
1
are compared with those
results of the proposed method (Table 5). The average value
of Column 3 in Table 5 is 1.00, while the average value of
Fig. 7(a). Two-story one-bay frame with
uniformly distributed loads.
Fig. 7(b). Two-story one-bay frame with
uniformly distributed loads.
Table 2.
Maximum Moments in Elastic Semi-Rigid
Frame Analysis (in.-kips)
(Two-story one-bay frame, Figure 6)
Element
No.
(1) (2) (3)
Second-Order
(Exact) Proposed (2) / (1)
1 1560 1696 1.09
2 1116 1037 0.93
3 1837 1847 1.01
4 1834 1839 1.00
5 1116 1037 0.93
6 1116 1037 0.93
Table 3.
Lateral Displacements at Windward
Beam-Column Joints (in.)
(Two-story one-bay frame, Figure 6)
Node
No.
Rigid Frame Semi-Rigid Frame
(1) (2) (3) (4) (5) (6)
Second-
Order
(Exact) Proposed (2) / (1)
Second-
Order
(Exact) Proposed (5) / (4)
3 1.16 1.21 1.04 2.02 1.85 0.92
5 1.73 1.82 1.05 3.26 2.98 0.91
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 136
Column 5 is 0.97. It can be seen that the Barakat method is
less conservative in this case.
6.4 Four-Story Two-Bay Frame with Uniformly
Distributed Loads
A four-story two-bay frame as shown in Figure 10 is investi-
gated here for the maximum column moments in both rigid
and semi-rigid frames. The semi-rigid connection of Jenkins
is utilized. The average value of Column 3 in Table 6 is 1.06,
while the average value of Column 6 in Table 6 is 1.01. The
proposed method represents reasonably well the second-or-
der effect for rigid and semi-rigid frames. The lateral dis-
placements at windward beam-column joints are summarized
in Table 7. The lateral displacements predicted by the pro-
posed method are less than those of the second-order elastic
semi-rigid frame analysis. However, the lateral displacements
predicted by the proposed method are larger than that of the
second-order elastic rigid frame analysis.
7. SUMMARY AND CONCLUSIONS
Several conclusions can be drawn from the present studies:
1. The moment-rotation relationships of semi-rigid con-
nections as represented by a simple tangent stiffness Fig. 8. Experimental III-17 connection curves (Barakat, 1991)
Table 4(a).
Maximum Moments in Elastic Semi-Rigid
Frame Analysis (in.-kips)
(Two-story one-bay frame, Figure 7a)
Element
No.
(1) (2) (3) (4) (5)
Second-
Order
(Exact) Proposed (2) / (1) Barakat (4) / (1)
1 257 220 0.86 201 0.78
2 547 588 1.07 576 1.05
3 526 552 1.05 557 1.06
4 813 818 1.00 811 0.99
5 1497 1431 0.96 1434 0.96
6 1497 1431 0.96 1434 0.96
7 940 922 0.98 923 0.98
8 940 922 0.98 923 0.98
Table 4(b).
Maximum Moments in Elastic Semi-Rigid
Frame Analysis (in.-kips)
(Two-story one-bay frame, Figure 7b)
Element
No.
(1) (2) (3) (4)
Linear-
Elastic
Rigid
Second-Order
Elastic
Semi-Rigid Proposed (3) / (2)
1 88 176 168 0.95
2 324 278 276 0.99
3 283 188 187 0.99
4 400 343 338 0.99
5 562 728 718 0.99
6 673 728 718 0.99
7 384 461 463 1.00
8 400 461 463 1.00
Table 4(c).
Lateral Displacements at Windward
Beam-Column Joints (in.)
(Two-story one-bay frame, Figure 7b)
Node
No.
Rigid Frame Semi-Rigid Frame
(1) (2) (3) (4)
Linear
Elastic
(Exact)
Second-
Order
(Exact) Proposed (3) / (2)
3 0.14 0.25 0.24 0.96
6 0.23 0.50 0.47 0.94
FOURTH QUARTER / 1993 137
expression are convenient and can lead to a close mo-
ment-rotation curve by numerical integration when
compared with test result. Note that only the tangent
stiffness is needed in an incremental nonlinear frame
analysis.
2. The proposed method gives close results to that of sec-
ond-order elastic analysis. It can handle both the uni-
formly distributed gravity loads and concentrated loads,
and predicts well the drift of unbraced frames.
3. All mean values of the normalized moment ratios are
found close to or slightly greater than one in the pro-
posed method. This shows that the proposed method is
more accurate when compared with that of the Barakat
method. The proposed method gives a reasonable pro-
cedure for estimating the approximate P- column mo-
ments for both rigid and semi-rigid frames.
4. The notional lateral loads calculation is relatively simple
and straightforward because the tedious determination
of the effective length factor, K, can be avoided. It is a
simple and practical method for semi-rigid frame design.
REFERENCES
1. Barakat, M. and Chen, W. F., Design Analysis of Semi-
Rigid Frames: Evaluation and Implementation, AISC,
Engineering Journal, 2nd Qtr., 1991, pp. 5564.
2. Yee, Y. L. and Melchers, R. E., Moment-Rotation Curves
for Bolted Connections, ASCE, J Struct. Eng., 112(3),
1986, pp. 615634
3. Jenkins, W. M., Tong, C. S. and Prescott, A. T., Moment-
Transmitting End-Plate Connections in Steel Construc-
Fig. 10. Four-story two-bay frame.
Fig. 9. Three-story one-bay frame with
uniformly distributed loads.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 138
tion, and a Proposed Basis for Flush End-Plate Design,
Struct. Engrg., 64A(5), 1986, pp. 121132.
4. Azizinamini, A., Bradburn, J. H., and Radziminski, J. B.,
Initial Stiffness of Semi-Rigid Steel Beam-to-Column
Connections, J. Construct. Steel Research 8, 1987, pp.
7190
5. Richard, R. M., Hsia, W. K. and Chmielowiec, M., Mo-
ment Rotation Curves for Double Framing Angles, Ma-
terials and Member Behavior, 1987, 107121.
6. Richard, R. M., Gillett, P. E., Kriegh, J. D. and Lewis, B.
A., The Analysis and Design of Single-Plate Framing
Connections, AISC, Engineering Journal, 2nd Qtr.,
1980, pp. 3852.
7. Frye, M. J. and Morris, G. A., Analysis of Flexibly
Connected Steel Frames, Can. J. Civ. Eng., 1975, 2(3),
pp. 280291.
8. Ang, K. M. and Morris, G. A., Analysis of Three-Dimen-
sional Frames with Flexible Beam-Column Connec-
tions, Can. J. Civ. Eng., 11, 1984, pp. 245254.
9. Romstad, K. M. and Subramanian, C. V., Analysis of
Frames with Partial Connection Rigidity, ASCE, J.
Struct. Div., 96(11), 1970, pp. 22832300.
10. Ackroyd, M. H., Simplified Frame Design of Type PR
Construction, AISC, Engineering Journal, 4th Qtr.,
1987, pp. 14146.
11. Goto, Y. and Chen, W. F., Second-Order Elastic Analysis
Table 5.
Maximum Moments in Elastic Semi-Rigid
Frame Analysis (in.-kips)
(Three-story one-bay frame, Figure 9)
Element
No.
(1) (2) (3) (4) (5)
Second-
Order
(Exact) Proposed (2) / (1) Barakat (4) / (1)
1 599 575 0.96 533 0.89
2 845 879 1.04 836 0.99
3 152 123 0.81 115 0.75
4 618 669 1.08 659 1.07
5 349 356 1.02 367 1.05
6 659 663 1.01 651 0.99
7 1181 1075 0.91 1076 0.91
8 1212 1386 1.14 1322 1.09
9 1082 1023 0.95 1025 0.95
10 1082 1148 1.06 1101 1.02
11 722 714 0.99 715 0.99
12 722 714 0.99 715 0.99
Table 6.
Maximum Column Moments in Elastic
Frame Analysis (in.-kips)
(Four-story two-bay frame, Figure 10)
Element
No.
Rigid Frame Semi-Rigid
(1) (2) (3) (4) (5) (6)
Second-
Order
(Exact) Proposed (2) / (1)
Second-
Order
(Exact) Proposed (5) / (4)
1 534 632 1.18 843 799 0.95
2 958 1066 1.12 1170 1182 1.01
3 1202 1296 1.08 1397 1398 1.00
4 455 421 0.93 291 322 1.11
5 656 729 1.11 596 698 1.17
6 1101 1142 1.04 1044 1061 1.02
7 615 603 0.98 559 558 0.99
8 473 525 1.11 542 562 1.04
9 1029 1061 1.03 996 1012 1.02
10 702 703 1.00 705 672 0.95
11 200 221 1.11 313 269 0.86
12 818 829 1.01 846 812 0.96
Table 7.
Lateral Displacments at Windward
Beam-column Joints (in.)
(Four-story two-bay frame, Figure 10)
Node
No.
Rigid Frame Semi-Rigid Frame
(1) (2) (3) (4) (5) (6)
Second-
Order
(Exact) Proposed (2) / (1)
Second-
Order
(Exact) Proposed (5) / (4)
4 0.27 0.30 1.11 0.40 0.37 0.93
9 0.66 0.73 1.11 1.07 0.95 0.89
14 0.94 1.04 1.11 1.61 1.40 0.87
19 1.11 1.23 1.11 1.95 1.68 0.86
FOURTH QUARTER / 1993 139
for Frame Design, ASCE, Journal of Structural Engi-
neering, Vol. 113, No. 7, 1987, pp. 15011519.
12. Chen, W. F. and Lui, E. M., Structural Stability: Theory
and Implementation, Elsevier, New York, 1987.
13. Chen, W. F. and Lui, E. M., Stability Design Criteria for
Steel Members and Frames in the United States, J. of
Constr. Steel Research, 5, Great Britain, 1985, pp. 3174.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 140
ABSTRACT
This paper is concerned with I-beam to box-column connec-
tions stiffened externally. A design method to determine the
dimensions of T-stiffeners is proposed. Connections of I-
beams and box-columns for a wide range of dimensions were
studied by the finite-element method and found to satisfy the
basic design criteria for a moment connection. The Ramberg-
Osgood function was used to curve-fit the moment-rotation
curves based on the geometric parameters of the connections
and the results are found to agree well with those from the
finite-element analyses. Finally, the design procedure and the
curve-fitting parameters were compared with the experimen-
tal results of a four-way connection tested to failure.
INTRODUCTION
It is a well known fact that the behavior of beams and columns
at their connection is one of the most important factors
considered in the analysis of steel frames. A vast number of
different types of connections are used, and the rigidity of
connections range from one that is extremely flexible, behav-
ing more like a pin joint, to one that is almost rigid. Re-
searchers have carried out studies on the effect of the semi-
rigid connection on frame behavior. A state-of-the-art paper
was presented by Jones, et al
1
on the analysis of frames with
semi-rigid joints. A modified stiffness matrix method, incor-
porating the partial rigidity of joints to find the elastic buck-
ling load of semi-rigid frames was presented by Yu and
Shanmugam.
2
Gerstle
3
noted that the effect of the connection
flexibility on frames can be two-fold: (a) connection rotation
contributes to the overall frame stability and (b) it affects the
distribution of internal forces and moments in the girders and
columns. The effects of connections on columns was consid-
ered by Nethercot and Chen
4
and Jones, et al
5
while Kato, et
al
6
carried out a study on the effect of joint flexibility due to
joint-panel shear deformation on frames. Barakat and Chen
7
used idealized connection models in the analysis of frames
and subsequently implemented the method on personal com-
puter.
Analytical models incorporating semi-rigid connections
will result in efficient design. In such design the frame mem-
bers will be utilized more efficiently, resulting in a lower cost.
Therefore, there is an important need to accurately determine
the moment-rotation (M-) characteristics of various types of
connections and to define them in a convenient way suitable
for incorporating them in frame analyses. Many researchers
have carried out both analytical and experimental investiga-
tions on the behavior of different types of connections. Vari-
ous types of models defining the M- relationship, ranging
from the simple linear model in the 1930s to the present day
complicated cubic B-spline curve-fitting model have been
reported by Jones, et al
1
The curve fitting technique was also
used by Attiogbe and Morris.
8
Experimental data was fitted
to the Richard-Abbott function while Ang and Morris
9
used
the Ramberg-Osgood function for the curve fitting process.
Due to the diversity of the behavior of the connections,
researchers
10,11
have proposed a classification system in an
attempt to present the behavior of connections consistently.
An alternative approach was to build up a data base for the
various types of connections
12,13
so that designers can obtain
the necessary data for their specific use.
However, most of the past work was carried out on connec-
tions between I-beams and I-columns. Limited work is avail-
able on the behavior of I-beam to box-column connec-
tions.
14,15,16
The authors have carried out an investigation on
such connections stiffened externally using the finite element
method.
17
Experimental and analytical results from tests car-
ried out on a series of specimens stiffened both internally as
well as externally have been reported.
18
It has been found that
the T-section provides an efficient external stiffener for the
connection. In this paper, a simple design procedure to deter-
mine the dimension of external T-stiffeners is proposed. A
curve fitting method using the Ramberg-Osgood function to
define the M- characteristic of such connections is also
discussed.
BASIC DESIGN PHILOSOPHY
The basic design criteria for rigid or moment connections
are:
19
1. sufficient strength
Lai-Choon Ting, Nandivaram E. Shanmugam and Seng-Lip Lee
are research assistant, associate professor and emeritus profes-
sor, respectively, of the Department of Civil Engineering, National
University of Singapore, Singapore.
Design of I-Beam to Box-Column Connections
Stiffened Externally
LAI-CHOON TING, NANDIVARAM E. SHANMUGAM AND SENG-LIP LEE
FOURTH QUARTER / 1993 141
2. sufficient rotation capacity
3. adequate stiffness
4. ease of erection and economical fabrication
Whether a particular connection satisfies the first three
conditions can be determined by observing its moment-rota-
tion curve. The last criterion is a matter of practical applica-
tion which has to take into account both the material cost of
the various components of the connection and the labor cost
in fabricating it.
Figure 1 shows four moment-rotation curves exhibiting
different characteristics of connections. Connection A is con-
sidered to be properly designed as it satisfies criteria (1) to
(3), i.e., it can attain sufficient strength in excess of the plastic
moment of the beam as well as having adequate stiffness and
rotation capacity before failure. Connection B, however, has
insufficient rotation capacity although it is adequate in terms
of stiffness and strength. As for connection C, it only has
rotation capacity but not stiffness and strength while connec-
tion D has neither sufficient strength nor rotation capacity. It
is the objective of this paper to present a design method for
I-beam to box-column connections stiffened externally with
T-stiffeners (Figure 2) which can exhibit the property of the
connection A in Figure 1.
CONNECTIONS WITH EXTERNAL T-STIFFENERS
When an I-beam frames into a box-column, the width of the
beam flange is normally less than the column width, as a result
the connection will be weak if it is not stiffened. In order to
achieve the conditions of an acceptable moment connection
(connection A in Figure 1), the traditional method is to stiffen
the connection by welding internal continuity plates at the
levels of the beam flanges inside the box-columns. This,
however, is a difficult and expensive process. The authors
have carried out a study
17,18
to investigate the possibility of
using external stiffeners in place of internal continuity plates
such that the basic design criteria are still satisfied.
The investigations showed that by using external T-stiffen-
ers, all the basic design criteria can be satisfied. Two series of
experiments have been carried out
18
on connection specimens
stiffened by internal continuity plates or by two different
types of external stiffeners namely angle and T-stiffeners. In
the first series, the specimens were subjected to a monotoni-
cally increasing load while in the second series, the specimens
were subjected to cyclic loads. The specimens consisted of
two 1.5 m long beams welded to opposite sides of a box-col-
umn of 1.0 m height. The dimensions of the external stiffeners
were designed based on a preliminary design method by using
finite element analysis. The length of the stiffeners were so
chosen that the normal stress distribution is uniform across
the stiffener. This would prevent any premature failure of the
stiffeners due to stress concentration. However, it was found
that this method gives rise to overdesign of the stiffeners and
an alternative method of design for the T-stiffener is, there-
fore, presented in this paper.
Due to the complexity of the connection involved, the finite
element method was used to analyze the connections.
MSC/NASTRAN,
20
which can carry out both material and
geometrical non-linear analyses was used to analyze all the
connections. Due to the symmetry of both the model and
loading, only a quarter of the model was analyzed and a
typical finite-element mesh is shown in Figure 3. The results
obtained are compared with those obtained from the experi-
ments.
T-STIFFENER DESIGN
Minimum stiffener length for stiffness
From the finite-element analyses, it was observed that a
minimum length for the T-stiffener is required to transfer the
forces from the beam flanges to the column webs effectively.
It was also found that for the stiffener to be effective, its web
thickness must be at least equal to half that of the beam-flange
Fig. 1. Moment-rotation curves. Fig. 2. Typical specimen with external T-stiffeners.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 142
thickness. Otherwise, the stiffener web may yield prema-
turely, resulting in a weak connection.
Figure 4 shows the load-deflection curves of a typical
connection with external T-stiffeners of various lengths.
These curves were obtained by the elasto-plastic finite-ele-
ment method using the MSC/NASTRAN package. The plas-
tic capacity of the beam, P
p
is also shown. The lengths were
defined by the angle as shown in the figure. The curves
corresponding to = 15 and 20 have sufficient stiffness and
strength while the other two curves are more flexible. The
curve corresponding to = 15 show marginal increase in
ultimate strength capacity over the curve with = 20. It was
thus decided to adopt = 20 as the design criterion because
this would result in a shorter stiffener length and hence more
economical design. The stiffener length l can thus be written
as
l = (B b) / (2 tan 20)
where
B = column width
b = beam-flange width
The other factor affecting the stiffness of the connection is
the stiffener-flange width which is connected to the edge of
the column web. This stiffener flange serves two purposes: 1)
it increases the moment of inertia of the beam cross-section
at the connection significantly, thus increasing the stiffness of
the connection, and 2) it transfers the stresses from the beam
to the column web more evenly, minimizing the possibility of
stress concentration.
Minimum Stiffener Length for Strength
The length of the stiffener, therefore, depends upon the ratio
of beam-flange width to column-flange width (b / B). When
this ratio reaches a value close to one, the stiffener length will
become so short that it will result in premature failure at the
stiffener web. For such cases, a check has to be made on the
minimum length based on the strength criteria of the stiffener
web. To determine the minimum length based on this type of
failure of the web, the following assumptions are made. The
moment developed at the connection should be at least equal
to the plastic moment capacity M
p
of the beam and it is carried
by the beam flanges such that T
p
= M
p
/ d
b
(Figure 5); the stress
distribution at failure on the beam flanges and stiffeners are
as shown in Figure 6(a) with stiffener flanges and stiffener
web between the flange and K-line reaching yield; the flange
forces, T
p
are transferred to the column webs through the
stiffeners as shown in Figure 6(b). It can be seen from Figure
6(b) that
T
p

2
T
1
+ T
2
(1)
where
T
1
= (A
f
+ A
w
) (2)
T
2
= lt
sw

y
(3)
A
f
= stiffener flange area
A
w
= area of stiffener web between the flange and K-line
l = stiffener length
t
sw
= thickness of stiffener web
Fig. 3. Typical finite element mesh.
Fig. 4. Load-deflection curves of specimen
with various stiffener lengths.
FOURTH QUARTER / 1993 143

y
=
yt
3

yt
= tensile strength
The stiffener length l can be calculated for Equations 13.
DESIGN PROCEDURE
Given the dimensions of a beam and a column, the following
simple procedure can thus be adopted to determine the suit-
able size and length of the T-stiffener required. From the
section table, an I-beam or T-section having web thickness
equal to at least half the beam-flange thickness is chosen for
the stiffener. Assuming = 20 (Figure 4), the stiffener length
is determined; the stiffener length based on the strength
criteria is calculated from Equations 13. The larger stiffener
length is finally chosen. All welds between the various com-
ponents at the connection are assumed to be full penetration
welds. Two examples based on the above design procedures
are shown in Appendix II.
LOAD-DEFLECTION CURVES FOR
TYPICAL SPECIMENS
Since there is no closed-form solution to define the behavior
of connections, the finite-element method has been used to
analyze these connections. This method has been shown to
predict the load-deflection characteristic of specimens with
reasonable accuracy. As such, it was decided to use this
method to test the validity of the design procedure proposed.
Two series of specimens were designed based on the above
procedure. One series consisted of specimens with two beams
framing into the box column on opposite sides while the other
series consists of connections with four beams framing into
the column on all four sides. The same design procedure was
used for both series, resulting in the same stiffener size for
connections between a particular beam and column dimen-
sions. Both the beam and box-column sections were obtained
from section tables.
22
Beams and column sizes were taken
such that the whole range of sections in the table can be
represented. The external T-stiffeners were then designed
accordingly and the specimens were analyzed using
MSC/NASTRAN.
Figure 7 shows some typical load-deflection curves of one
of the specimens. For comparison, the results obtained using
the simple elastic-plastic method are also plotted. It can be
seen from the figure that the connections are able to develop
strength well in excess of the plastic capacity of the beams.
In addition, the initial stiffness of the connections satisfied the
basic criteria for a moment connection. The slight difference
in the initial stiffness between the two- and four-way connec-
tions is expected since the column web for the two-way
connection is unrestrained in one direction while that of the
four-way connection is restrained all round.
MOMENT-ROTATION PREDICTION BY
CURVE-FITTING TECHNIQUE
It is commonly known that the behavior of connections
between beams and columns is one of the most uncertain
parameters in the design of frames at present. Analyses are
carried out assuming the connections to be either fixed or
pinned but in practice it is never the case. An accurate mo-
ment-rotation (M-) characteristic for different types of con-
nection is therefore essential if any work is to be carried out
to incorporate the semi-rigid nature of these connections.
Many researchers have tried to standardize the M- relation-
ships for various types of connections so that they can be
incorporated into the computer programs during the analysis
of frames. One of the most common method of standardizing
the M- curves is by curve-fitting the available test data for
the different types of connections.
Fig. 5. Internal forces at connection under symmetrical load. Fig. 6. Stress distribution at failure.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 144
STANDARDIZED MOMENT-ROTATION
FUNCTIONS
In the present study, a series of 15 M- curves has been
generated for each of the two cases, namely, two-way and
four-way connections using the finite-element method. These
specimens cover the combination of full range of beams and
box columns available in the section table.
22
The Ramberg-
Osgood function was used to curve-fit the data. This function
can be expressed in terms of moment, M, and rotation, , as
follows:


M
M
o

1 +

'

M
M
o

(n 1)
1
1
]
(4)
where M
o
,
o
, and n are the independent parameters of the
function. M
o
and
o
are the reference moment and rotation
respectively, while n defines the sharpness of the curve. These
independent parameters can be expressed in terms of the
geometric properties of the connection as follows:
M
o


i 1
m
p
i
a
i
(5)


i 1
m
p
i
b
i
(6)
n


i 1
m
p
i
c
i
(7)
where p
i
represents the ith geometric parameter of the con-
nection and a
i
, b
i
, and c
i
are the exponents that indicate the
effect of the ith geometric parameter; m is the number of
geometric parameters considered. Taking the logarithms of
both sides of the above equations, the Ramberg-Osgood
parameters can be expressed as:
log M
o
a
1
log p
1
+ a
2
log p
2
+...+ a
m
log p
m
(8)
log
o
b
1
log p
1
+ b
2
log p
2
+...+ b
m
log p
m
(9)
log n c
1
log p
1
+ c
2
log p
2
+...+ c
m
log p
m
(10)
Multiple linear regression analysis was then carried out to
determine the coefficients a, b, and c.
A total of six terms, representing the various geometries of
the connection, has been used to determine the coefficients,
and the relationships thus obtained are given as follows:
M
o

M

B
t
c
1
1
]
0.484

b
B
1
1
]
0.484

h
B
1
1
]
1.085
[
d
b]
2.738

t
sf
t
c
1
1
]
0.640

t
sw
t
bf
1
1
]
0.899
(11)

B
t
c
1
1
]
0.928

b
B
1
1
]
1.658

h
B
1
1
]
1.377
[
d
b]
0.887

t
sf
t
c
1
1
]
0.236

t
sw
t
bf
1
1
]
0.388
(12)
n
n

B
t
c
1
1
]
1.905

b
B
1
1
]
0.467

h
B
1
1
]
0.899
[
d
b]
0.222

t
sf
t
c
1
1
]
1.136

t
sw
t
bf
1
1
]
0.254
(13)
where
B = column width
b = beam flange width
t
c
= column thickness
h = stiffener flange width
d
b
= beam depth
t
sf
= stiffener flange thickness
t
sw
= stiffener web thickness
t
bf
= beam flange thickness

M
= 5.395 10
6
and 5.935 10
6
,

= 0.0324 and 0.0308


and
n
= 0.019 and 0.0285 for the two-way and four-way
connections respectively. All dimensions are in millimeters.
Units for M and M
o
are in kNm, and
o
are in radians ,while
the geometrical parameters were measured in millimeters.
Figure 8 shows the comparison between results obtained
from the Ramberg-Osgood function by using the standardized
connection parameters and the corresponding results from the
finite element analysis. Curves for typical specimens (Exam-
ple 1 in Appendix II) are shown for both the two-way and
four-way connections. It can be seen that the correlation
between the curves is very good. Similar observation has been
made for all the other specimens.
Figures 9(a) and (b) show plots of normalized moment-ro-
tation relationships for all the 15 specimens of two-way and
four-way connections, respectively. It can be seen that almost
all the curves in each case lie very close, except for two
specimens. These two specimens, which are the same for both
cases, consist of specimens with beam and column of extreme
sizes obtained from the section table i.e., one specimen con-
sists of the smallest column and very small beams while the
other specimen consists of the largest column size with very
large beams. It is suggested that a single curve, as shown in
Fig. 7. Typical load-deflection curves for
three-way and four-way connections.
FOURTH QUARTER / 1993 145
the respective figures, can be used for the prediction of the
moment-rotation characteristic of the connections between
all beams and columns of practical dimensions.
EXPERIMENTAL VERIFICATION
An experimental investigation was carried out to study the
behavior of a connection in which four I-beams frame into a
box column. This is the typical connection which could occur
at an interior column of a building. The connection was tested
to failure, and strain and displacement measurements were
made to obtain the stress distribution and moment-rotation
characteristics. Details of the experimental program together
with the results have been reported elsewhere.
23
For compari-
son, results from one of the specimens are shown here.
Figure 10 is typical specimen which was designed in
accordance with the proposed method. It was supported at the
ends of the four beams and subjected to a load applied
vertically on the column until failure. Figures 11 and 12 show
the load-deflection and the dimensionless moment-rotation
curves, respectively. It can be seen that the initial stiffness of
the connection is good and the ultimate strength capacity
exceeds the plastic capacity of the beam. The large rotation
capacity of the specimen also indicates that the connection is
ductile. Good correlation is observed between the results
from the experiment and the finite element method. Also, it
can be seen that the fitted Ramberg-Osgood curve using the
standardized moment-rotation function agrees with the ex-
perimental results.
CONCLUSIONS
A simple design procedure has been proposed to determine
the size of the T-section to be used as the external stiffener for
Fig. 8. Normalized moment-rotation
curves for typical speciments. Fig. 9. Normalized moment-rotation curves for 15 specimens.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 146
an I-beam to box-column connection. The design method is
applicable for both the two-way and four-way connections.
Results from connections, formed by a wide range of I-beams
and box columns obtained from section tables, show the
accuracy of the proposed design method.
It is common practice in modern design codes to propose
a capacity reduction factor for ultimate-strength design. A
similar factor may be adopted for the calculation of T
1
and
T
2
to account for various uncertainties in the connection
design and performance. A value of 0.9 given in the AISC-
LRFD Specification for both shear and tension yield limit
states may be used in this case also.
A curve-fitting procedure using the Ramberg-Osgood
function has been used to obtain the moment rotation rela-
tionship for the connections. The independent parameters of
the function was expressed in terms of the geometrical prop-
erty of the connection. The curves obtained by using the fitted
parameters compare well with those obtained from the finite-
element analysis for all the specimens considered. Results
indicate that a single moment-rotation curve can be used for
the design of connections formed by a wide range of beam
and column sections. Finally, both the design procedure and
the Ramberg-Osgood function obtained have been found to
agree well with experimental results. The results presented
are with reference to two-way and four-way connections.
Further research is in progress to study other types of configu-
rations and to investigate the suitability of this connection for
seismic design.
NOTATION
B column width
M moment imposed on the connection
M
o
reference moment
M
p
plastic moment of beam
T
p
beam-flange force corresponding to plastic
moment of beam
a
i
, b
i
, c
i
exponents indicating effect of ith geometric
parameter
b beam flange width
d
b
beam depth
h stiffener flange width
l stiffener length
m number of geometric parameters considered
n sharpness of Ramberg-Osgood curve
p
i
ith geometric parameter of connection
t
bf
beam flange thickness
t
c
column wall thickness
t
sf
stiffener flange thickness
t
sw
stiffener web thickness
rotation at the connection

o
reference rotation
ACKNOWLEDGEMENT
The investigation presented in this paper is part of a program
of research on box columns being carried out in the Depart-
ment of Civil Engineering at the National University of
Singapore. The work is funded by research grant RP94/85
made available by the National University of Singapore.
APPENDIX IREFERENCES
1. Jones, W. S., Kirby, P. A., and Nethercot, D. A., The
Analysis of Frames with Semi-Rigid ConnectionsA
State-of-the-Art Report, Journal of Constructional Steel
Research, Vol. 3, No. 2, 1983, pp. 213.
2. Yu, C. H. and Shanmugam, N. E., Stability of Frames
with Semi-Rigid Joints, Computers and Structures, Vol.
23, No. 5, 1986, pp. 639648.
3. Gerstle, K. H., Effect of Connection on Frames, Journal
of Constructional Steel Research, Vol. 10, 1988, pp. 241
267.
4. Nethercot, D. A., and Chen, W. F., Effects of Connections
on Columns, Journal of Constructional Steel Research,
Vol. 10, 1988, pp. 201239.
5. Jones, W. S., Kirby, P. A., and Nethercot, D. A., Effect of
Fig. 10. 4-way connection test specimen. Fig. 11. Load-deflection curve for 4-way test specimen.
FOURTH QUARTER / 1993 147
Semi-Rigid Connections on Steel Column Strength,
Journal of Constructional Steel Research, Vol. 1, 1980,
pp. 3846.
6. Kato, B., Chen, W. F., and Nakao, M., Effects of Joint-
Panel Shear Deformation on Frames, Journal of Con-
structional Steel Research, Vol. 10, 1988, pp. 269320.
7. Barakat, M. and Chen, W. F., Design Analysis of Semi-
Rigid Frames: Evaluation and Implementation, Engi-
neering Journal, AISC, 2nd Quarter, 1991, pp. 5564.
8. Attiogbe, E. and Morris, G., Moment-Rotation Func-
tions for Steel Connections, Journal of Structural Engi-
neering Division, ASCE, Vol. 117, ST6, 1991, pp. 1703
1718.
9. Ang. K. M. and Morris, G. A., Analysis of Three-Dimen-
sional Frames with Flexible Beam-Column Connec-
tions, Canadian Journal of Civil Engineering, Vol. 11
No. 2, 1984, pp. 245254.
10. Bjorhovde, R., Colson, A., and Brozzetti, J., Classifica-
tion System for Beam-to-Column Connections, Journal
of Structural Division, ASCE, Vol. 116, No. 11, Nov. l990,
pp. 30593076.
11. Maquoi, R., Semi-Rigid Joints: from Research to Design
Practice, International Conference on Steel and Alu-
minium Structures, Singapore, 2224 May, 1991, pp.
3243.
12. Chen, W. F. and Kishi, N., Semi-Rigid Steel Beam-to-
Column Connections: Data Base and Modelling, Jour-
nal of Structural Division, ASCE, Vol. 107, ST 9, 1981,
pp. 105119.
13. Nethercot, D. A., Utilization of Experimentally Ob-
tained Connection Data in Assessing the Performance of
Steel Frames, Connection Flexibility and Steel Frames,
W. F. Chen, Ed., 1985, pp. 1337.
14. Chen, S. J. and Lin, H. Y., Experimental Study of Steel
I-Beam to Box-Column Moment Connections, 4th Inter-
national Conference on Steel Structures and Space
Frames, Feb. 1516, 1990, Singapore, pp. 4147.
15. Dawe, J. L. and Grondin, G. Y., W-Shape Beam to RHS
Column Connections, Canadian Journal of Civil Engi-
neering, Vol. 17, Oct., 1990, pp. 788797.
16. White, R. N. and Fang, P. J., Framing Connections for
Square Structural Tubing, AISC National Engineering
Conference, Memphis, Tenn, April, 1965, pp. 74102.
17. Ting, L. C., Shanmugam, N. E., and Lee, S. L., Box-Col-
umn to I-Beam Connections Stiffened Externally, Jour-
nal of Constructional Steel Research, Vol. 18, No. 3, 1991,
pp. 209226.
18. Ting, L. C., Shanmugam, N. E., and Lee, S. L, Externally
Stiffened Steel Beam to Box-Column Connections, Pro-
ceedings, International Conference on Steel and Alu-
minium Structures, Singapore, May, 1991.
19. Shanmugam, N. E., Ting, L. C., and Lee, S. L., Behavior
of I-Beam to Box-Column Connections Stiffened Exter-
nally and Subjected to Fluctuating Loads, Journal of
Constructional Steel Research, Vol. 20, No. 2, 1991, pp.
129148.
20. Chen, W. F. and Lui, E. M., Static Flange Moment
Connections, Journal of Constructional Steel Research,
Vol. 10, 1988, pp. 3888.
21. MSC/NASTRAN Application Manual, Volumes I and II,
The MacNeal-Schwendler Corporation, April, 1983.
22. Steelwork Design Guide to BS5950: Part 1, Vol. 1, Section
Properties, Member Capacities, 2nd ed., The Steel Con-
struction Institute, 1985.
23. Lee, S. L., Ting, L. C., and Shanmugam, N. E., Use of
External T-Stiffeners in Box-Column to I-Beam, Journal
of Constructional Steel Research, Vol. 26, Nos. 2&3,
1993, pp. 7798.
APPENDIX IIDESIGN EXAMPLES
Design examples to determine the size of the T-section which
can be used as the external stiffener for I-beam to box-column
connections. Example 1 illustrates the case when the stiffener
length is governed by the angle 2 while Example 2
illustrates the case when the length is governed by the shear
capacity of the stiffener web.
Example 1
Column size: 20020016 mm
Beam size: 30516540 kg/m
Grade 43 steel is assumed for column, beam, and stiffeners.
1. Choose a T- or I-section with a web thickness of at least
half the beam-flange thickness.
Try T-section 10210212 kg/m
2. For = 20, stiffener length l = (200 165) / (2 tan
20) = 50 mm (2 in.)
3. Check minimum stiffener length based on strength criteria.
Plastic moment capacity of the beam, Fig. 12. Moment-rotation curve for four-way test specimen.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 148
M
p
= 275 624,000 Nmm
= 172 kNm (126.9 kip-ft)
T
p
= 172 / 0.304 = 584 kN (131.3 kips)
T
1
= (101.6 9.3 + 7.6 5.2)0.275 = 281 kN (63.2 kips)
T
2
= 0.5T
p
T
1
= 11 kN (2.47 kips)
The allowable shear stress according to von Mises yield
criteria,

y
= 275 / 1.732 = 159 N/mm
2
(23.1 ksi)
From Equation 1
T
p
/ 2 = T
1
+ T
2
Therefore, the minimum length is
l = 11,000 / (159 5.2) = 14 mm (0.55 in.)
4. From Steps 2 and 3, the longer length is chosen.
Therefore, use section 10210212 kg/m of length 50 mm
(2 in.) as the external stiffener.
Example 2
Column size: 25025016 mm
Beam size: 45719167 kg/m
Grade 43 steel is assumed for column, beam, and stiffeners.
1. Choose a T- or I-section with a web thickness of at least
half the beam-flange thickness. Try T-section
10212714 kg/m
2. For = 20, stiffener length l = (250 191) / (2 tan
20) = 83 mm (3.27 in.)
3. Check minimum stiffener length based on strength cri-
teria. Plastic moment capacity of the beam,
M
p
= 275 1,470,000 Nmm
= 404 kNm (298 kip-ft)
T
p
= 404 / 0.441 = 916 kN (205.9 kips)
T
1
= (102.1 10 + 7.6 6.4) = 294 kN (66.1 kips)
T
2
= 0.5T
p
T
1
= 164 kN (36.9 kips)
The allowable shear stress according to von Mises yield
criteria,

y
= 275 / 1.732 = 159 N/mm
2
(23.1 ksi)
Therefore, the minimum length
l = 164,000 / (159 6.4) = 161 mm (6.34 in.)
4. From Steps 2 and 3, the longer length is chosen.
Therefore, use section 10212714 kg/m of length 161 mm
(6.34 in.) as the external stiffener.
FOURTH QUARTER / 1993 149
The author has provided useful suggestions for this particular
design problem based on yield-line analysis. The purpose of
this discussion is to clarify some aspects of these solutions
and to re-organize them in concise decision form for ease of
calculations.
While the AISC Manuals have only exclusively addressed
base plates for wide-flange column shapes, the AISC Design
Guide No. 1 Column Base Plates (Ref. 3) does briefly cover
tubular and pipe columns. It suggests that the usual cantilever
plate model employed under wide-flange columns can be
extended to such closed sections: the critical overhang dimen-
sion (m or n) for determining plate thickness becomes 0.95
times the outside column dimension for rectangular tubes and
0.80 times the outside dimension for round pipes.
Also, similar to base plates with wide-flange columns, an
important consideration is the limiting case when the column
approaches the size of its base plate. For this so-called small
plate condition, the cantilever overhang distance m can become
rather short and almost zero, thereby rendering this simple model
useless for design. The proposed yield-line Equation 2 possesses
this same characteristic since for R / D = 1, the required thickness
reduces to zero. Fortunately, Equation 1 does provide a rational
design answer for R / D = 1.
While Equations 1 and 2 in combination offer valid and
complementary design solutions, the presented application of
lightly loaded Equation 3 in this context is confusing. The
paper states that Equation 3 is a special case of Equation 1.
The required plate thickness is limited to no more than given
by Equation 1 and, finally, the greater of Equations 1 or 3 and
2. Based on this rationale, it is never necessary to check
Equation 3. In order to parallel the logic of the revised AISC
small-column base-plate procedure for wide-flange shapes, it
appears that the real intent should be for the required thick-
ness to be the greater of:
a. the lesser of Equations 1 and 3 or, conservatively,
Equation 1
b. Equation 2
The authors Equations 1 and 2 may be interpreted such
that the former applies to small plates (R / D near 1.0)
whereas the latter covers the larger plates (smaller R / D).
Use of Equation 3, in my opinion, is optional in conjunction
with Equation 1 for lightly loaded conditions. The recom-
mended practical limit of R / D 0.5 gives an actual design
range of 0.5 R / D 1.0. One may easily compute that the
intersection of Equations 1 and 2 occurs at about R / D = 0.7,
hence, the following general design criteria can be formulated
to minimize calculations:
If 0.5 R / D 0.7, (large base plate case) use Equation 2.
.
If 0.7 < R / D 1.0, (small base plate case) use the lesser
of Equations 1 and 3, or, conservatively, Equation 1.
In the original example, R / D = 0.61 < 0.7, and Equation
2 governs, as expected.
As alluded to previously, the cantilever bending model
could also be utilized for the large base-plate case. Because
its solution is slightly more conservative than given by Equa-
tion 2, the R / D = 0.7 limit should be increased to 0.8 for its
range of applicability. Applying this procedure to the authors
example problem results in:
R / D 0.61 < 0.8 o.k.
m D 0.8R 3.5 0.8 (2.13) 1.796 in.
f
p

P
4D
2

12
4(3.5)
2
0.245 ksi
t 2m

f
p
F
y
2(1.796)

.245
36
0.296 in.
This solution requires an extra one-sixteenth base-plate thick-
ness compared to the yield-line based Equation 2. However,
it can also serve to demonstrate the reasonableness of both
methods.
Easy conversions can be made for LRFD design equivalent
to the proposed ASD Equations 1, 2, and 3:
1. replace M by M
p
0.9t
2
F
y
/ 4
2. replace P by P
u
(factored loads) in f
p
3. solve appropriate work balance expressions for required
thickness
DISCUSSION
Design of Pipe Column Base Plates Under Gravity Load
Paper by THOMAS SPUTO
(2nd Quarter, 1993)
Discussion by Nestor R. Iwankiw
Nestor R. Iwankiw is Director of Research & Codes, AISC,
Chicago, IL.
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 150
LRFD:
t
R
D
P
u
/ (2.7F
y
) (1)
t

0.74P
u

F
y

2 3

R
D
_

,
+

R
D
_

,
3
1
1
]
(2)
R
c
R
o
2
P
u
/ F
p
(3)
t R

F
p

2.7F
y

1 3

R
c
R
_

,
+ 2

R
c
R
_

,
3
1
1
]
In summary, this paper contributes new and useful design
criteria for column base plates under gravity loads which,
with additional reflection, can be further simplified and gen-
eralized for applications.
CLOSURE BY THOMAS SPUTO
The discussor has provided some interesting suggestions for
simplifying the content of this paper. His clarification of the
applicability of each design equation is especially welcome.
The author thanks him for his interest in this topic.
ERRATA
1. should be included as multiplier in the final external
work expression above Equation 2.
2. Equation 3 and W
e
equation above it: change exponent
in last term on R
c
from 2 to 3.
3. Example
t by Equation 2, last term should be changed from
(2.13)
2
/ 3.15 to (2.13)
3
/ 3.5
4. Page 42, upper left, external work expression should
read
f
p

D
2

D
2
3
(D
2
R
2
)

R
D
_

,
R
2

R
D
_

,
+
R
2
3

R
D
_

,
1
1
]
f
p

2D
2
3
RD +
R
3
3D
1
1
]
FOURTH QUARTER / 1993 151
CORRECTION
Shear Tab Design Tables
ASD/LRFD Volume IIConnections
The following are corrected tables for pages C-11 and C-12 of the AISC Manual of Steel Construction Volume IIConnections.
Single-Plate Shear Connections
Rigid SupportStandard Holes
Allowable loads in kips
n = 2 L = 6
Plate
Thickness, t
in.
Bolt Size, in.
3

4
7

8 1
Load Weld Load Weld Load Weld
1

4
5

16
3

8
7

16
1

2
9

16
10.9
10.9
10.9
10.9

16
1

4
5

16
3

14.9
14.9
14.9
14.9
14.9

16
1

4
5

16
3

8
3

16.9
19.5
19.5
19.5
19.5
19.5
3

16
1

4
5

16
3

8
3

8
7

16
1

4
5

16
3

8
7

16
1

2
9

16
14.6
14.6
14.6
14.6

16
1

4
5

16
3

17.9
19.9
19.9
19.9
19.9

16
1

4
5

16
3

8
3

16.9
21.1
25.3
25.9
25.9
25.9
3

16
1

4
5

16
3

8
3

8
7

16
Single-Plate Shear Connections
Rigid SupportStandard Holes
Design loads in kips
n = 2 L = 6
Plate
Thickness, t
in.
Bolt Size, in.
3

4
7

8 1
Load Weld Load Weld Load Weld
1

4
5

16
3

8
7

16
1

2
9

16
18.3
18.3
18.3
18.3

16
1

4
5

16
3

24.9
24.9
24.9
24.9
24.9

16
1

4
5

16
3

8
3

25.3
31.6
32.5
32.5
32.5
32.5
3

16
1

4
5

16
3

8
3

8
7

16
1
4
5

16
3

8
7

16
1

2
9

16
22.9
22.9
22.9
22.9

3
16
1

4
5

16
3

26.9
31.1
31.1
31.1
31.1

3
16
1

4
5

16
3

8
3

25.3
31.6
37.9
40.7
40.7
40.7
3
16
1

4
5

16
3

8
3

8
7

16
Single-Plate Shear Connections
Rigid SupportStandard Holes
Allowable loads in kips
n = 3 L = 9
Plate
Thickness, t
in.
Bolt Size, in.
3

4
7

8 1
Load Weld Load Weld Load Weld
1

4
5

16
3

8
7

16
1

2
9

16
24.1
24.1
24.1
24.1

16
1

4
5

16
3

26.9
32.8
32.8
32.8
32.8

16
1

4
5

16
3

8
3

25.3
31.6
37.9
42.9
42.9
42.9
3

16
1

4
5

16
3

8
3

8
7

16
1

4
5

16
3

8
7

16
1

2
9

16
28.5
32.2
32.2
32.2

16
1

4
5

16
3

26.9
33.6
40.4
43.8
43.8

16
1

4
5

16
3

8
3

25.3
31.6
37.9
44.2
50.6
56.9
3

16
1

4
5

16
3

8
3

8
7

16
Single-Plate Shear Connections
Rigid SupportStandard Holes
Design loads in kips
n = 3 L = 9
Plate
Thickness, t
in.
Bolt Size, in.
3

4
7

8 1
Load Weld Load Weld Load Weld
1

4
5

16
3

8
7

16
1

2
9

16
40.3
40.3
40.3
40.3

16
1

4
5

16
3

40.4
50.5
54.9
54.9
54.9

16
1

4
5

16
3

8
3

37.9
47.4
56.9
66.4
71.7
71.7
3

16
1

4
5

16
3

8
3

8
7

16
1

4
5

16
3

8
7

16
1

2
9

16
42.8
50.4
50.4
50.4

16
1

4
5

16
3

40.4
50.5
60.6
68.6
68.6

16
1

4
5

16
3

8
3

37.9
47.4
56.9
66.4
75.9
85.3
3

16
1

4
5

16
3

8
3

8
7

16
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 152
COMPOSITE DESIGN
Composite Girders with Partial Restraints: A New
Approach Wexler, Neil . . . . . . . . . . . . . . . . . . . . . 68
Strength of Shear Studs in Steel Deck on Composite
Beams and Joists Easterling, W. Samuel, David R.
Gibbings, and Thomas M. Murray . . . . . . . . . . . . . . 44
BEAMS
The Warping Contstant for the W-Section with a
Channel Cap Lue, Tony and Duane S. Ellifritt . . . 31
COLUMNS
Design of Pipe Column Base Plates Under Gravity
Load Sputo, Thomas . . . . . . . . . . . . . . . . . . . . . . . 41
DiscussionDesign of Pipe Column Base Plates Under
Gravity Load Sputo, Thomas and Nestor R.
Iwankiw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
CONNECTIONS
A Tentative Design Guideline for a New Steel Beam
Connection Detail to Composite Tube Columns
Azizinamini, Atorod and Bangalore Prakash . . . . . 108
CorrectionFast Check for Block Shear Burgett,
Lewis B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Design Aid of Semi-Rigid Connections for Frame
Analysis Kishi, N., W. F. Chen, Y. Goto, and K. G.
Matsuoka . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Design of I-Beam to Box-Column Connections Stiffened
Externally Ting, Lai-Choon, Nandivaram E.
Shanmugam and Seng-Lip Lee . . . . . . . . . . . . . . . . 141
The Economic Impact of Overspecifying Simple
Connections Carter, Charles J. and Louis F.
Geschwindner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
FRAMES
Composite Girders with Partial Restraints: A New
Approach Wexler, Neil . . . . . . . . . . . . . . . . . . . . . 68
Composite Semi-Rigid Construction Leon,
Roberto T.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
CorrectionSimple Equations for Effective Length
Factors Dumonteil, Pierre. . . . . . . . . . . . . . . . . . . 38
Design Aid of Semi-Rigid Connections for Frame
Analysis Kishi, N., W. F. Chen, Y. Goto, and K. G.
Matsuoka . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
DiscussionSimple Equations for Effective Length
Factors Moore, William E. II. . . . . . . . . . . . . . . . . 37
GIRDERS
Composite Girders with Partial Restraints: A New
Approach Wexler, Neil. . . . . . . . . . . . . . . . . . . . . . 68
LOAD AND RESISTANCE FACTOR DESIGN
CorrectionASD/LRFD Volume IIConnections
(Shear Tab Design Tables) Carter, Charles J. and
Nestor Iwankiw . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
EARTHQUAKE DESIGN
Earthquakes: Steel Structures Performance and Design
Code Developments Marsh, James W. . . . . . . . . . 56
SERVICEABILITY
Serviceability Limit States Under Wind Load Griffis,
Lawrence G. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
TENSION
Shear Lag Effects in Steel Tension Members
Easterling, W. Samuel and Lisa Gonzalez Giroux. . . 77
SINGLE-ANGLE
Design Strength of Concentrically Loaded Single-Angle
Struts Zureick, A. . . . . . . . . . . . . . . . . . . . . . . . . . . 17
METRIC
SI Units for Structural Steel Design American
Institute of Steel Construction . . . . . . . . . . . . . . . . . . 66
VIBRATION
Design Criterion for Vibrations due to Walking Allen,
D. E. and Thomas M. Murray . . . . . . . . . . . . . . . . . 117
Annual Index
First Quarter 139 Third Quarter 77115
Second Quarter 4175 Fourth Quarter 117152
SUBJECT INDEX
ENGINEERING JOURNAL / AMERICAN INSTITUTE OF STEEL CONSTRUCTION 153
AUTHOR INDEX
Allen, D. E. and Thomas M. Murray
Design Criterion for Vibrations due to Walking . . . . . 117
American Institute of Steel Construction
SI Units for Structural Steel Design . . . . . . . . . . . . . . . 66
Azizinamini, Atorod and Bangalore Prakash
A Tentative Design Guideline for a New Steel Beam
Connection Detail to Composite Tube Columns. . . 108
Burgett, Lewis B.
CorrectionFast Check for Block Shear . . . . . . . . . . . 39
Carter, Charles J. and Louis F. Geschwindner
The Economic Impact of Overspecifying Simple
Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Carter, Charles J. and Nestor R. Iwankiw
CorrectionASD/LRFD Volume IIConnections
(Shear Tab Design Tables) . . . . . . . . . . . . . . . . . . . . 150
Chen, W. F.
See Kishi, N. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Dumonteil, Pierre
CorrectionSimple Equations for Effective Length
Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Easterling, W. Samuel and Lisa Gonzalez Giroux
Shear Lag Effects in Steel Tension Members . . . . . . . . 77
Easterling, W. Samuel, David R. Gibbings, and Thomas M.
Murray
Strength of Shear Studs in Steel Deck on Composite
Beams and Joists . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
Ellifritt, Duane
See Lue, Tony. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Geschwindner, Louis F.
See Carter, Charles J. . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Gibbings, David R.
See Easterling, W. Samuel . . . . . . . . . . . . . . . . . . . . . . . 44
Giroux, Lisa Gonzalez
See Easterling, W. Samuel . . . . . . . . . . . . . . . . . . . . . . . 77
Goto, Y
See Kishi, N. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Griffis, Lawrence G.
Serviceability Limit States Under Wind Load . . . . . . . . 1
Iwankiw, Nestor R.
See Sputo, Thomas. . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
Iwankiw, Nestor R.
See Carter, Charles J. . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Kishi, N., W. F. Chen, Y. Goto, and K. G. Matsuoka
Design Aid of Semi-Rigid Connections for Frame
Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Lee, Seng-Lip
See Ting, Lai-Choon . . . . . . . . . . . . . . . . . . . . . . . . . 141
Leon, Roberto T.
Composite Semi-Rigid Construction. . . . . . . . . . . . . 130
Lue, Tony and Duane S. Ellifritt
The Warping Contstant for the W-Section with a
Channel Cap. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Marsh, James W.
Earthquakes: Steel Structures Performance and Design
Code Developments . . . . . . . . . . . . . . . . . . . . . . . . . 56
Matsuoka, K. G.
See Kishi, N.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Moore, William E. II
DiscussionSimple Equations for Effective Length
Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Murray, Thomas M.
See Easterling, W. Samuel . . . . . . . . . . . . . . . . . . . . . . 44
Murray, Thomas M.
See Allen, D.E. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Prakash, Bangalore
See Azizinamini, Atorod . . . . . . . . . . . . . . . . . . . . . . 108
Shanmugam, Nandivaram E.
See Ting, Lai-Choon . . . . . . . . . . . . . . . . . . . . . . . . . 141
Sputo, Thomas
Design of Pipe Column Base Plates Under Gravity
Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Sputo, Thomas
DiscussionDesign of Pipe Column Base Plates Under
Gravity Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Ting, Lai-Choon, Nandivaram E. Shanmugam and
Seng-Lip Lee
Design of I-Beam to Box-Column Connections Stiffened
Externally . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Wexler, Neil
Composite Girders with Partial Restraints: A New
Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
Zureick, A.
Design Strength of Concentrically Loaded Single-Angle
Struts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
FOURTH QUARTER / 1993 154

También podría gustarte