You are on page 1of 24

32

Structural Bracing

32.1 Introduction .................................................. 32-1


32.2 Types of Bracing............................................. 32-1
32.3 Background ................................................... 32-2
Member Out-of-Straightness and Brace Stiffness  Member
Shortening  Member Inelasticity  Brace Connection
Stiffness
32.4 Safety Factors................................................. 32-4
32.5 Column or Frame Bracing ................................ 32-4
Column Buckling and Design Philosophy  Relative Systems 
Nodal Systems  Continuous Systems  Lean-On Systems 
Torsional Bracing
Brian Chen
Wiss, Janney, Elstner 32.6 Beam Bracing................................................. 32-12
Associates, Inc., Inflection Points as Brace Points  Lateral Bracing  Torsional
Irving, TX Bracing
32.7 Faulty Details ................................................. 32-18
Joseph Yura Nomenclature ........................................................ 32-21
Department of Civil Engineering,
University of Texas, References ............................................................. 32-22
Austin, TX Further Reading ..................................................... 32-22

32.1 Introduction
This chapter presents an overview of aspects related to the design of structural bracing used in beams,
columns, and frame structures and is intended for practicing civil and structural engineers. Many of the
design guidelines presented were incorporated into the 2002 Load and Resistance Factor Design Manual
published by the American Institute of Steel Construction (AISC). The intended focus is on simplicity
and ease of implementation over exact formulations. The basis for the design formulations along with
a classification system for bracing systems is first presented. Design formulations are presented with
illustrative numerical examples. Finally, common faulty bracing details are presented.

32.2 Types of Bracing


Bracing used in structural systems generally serve two primary functions. They resist secondary loads on
structures (e.g., wind bracing) and increase the strength of individual members by resisting deformation
in the weakest direction [1]. For the latter case, structural bracing forces higher modes of deformation by
providing resistance to lateral and/or rotational displacement. This is achieved through axial, shear, and/
or flexural deformations of the bracing member. Diaphragms, for instance, provide restraint through
their shear stiffness while diagonal cross-bracing relies on axial stiffness.
Bracing systems used to control instability fall into four general classifications: relative, nodal, con-
tinuous, or lean-on. Common configurations of each type are shown in Figure 32.1. Relative bracing

0-8493-1569-7/05/$0.00+$1.50
# 2005 by CRC Press 32-1

Copyright 2005 by CRC Press


32-2 Handbook of Structural Engineering

(a) (b)
Comp Cross
flange frames

Diaphragms
Brace

(c) (d)

Metal
deck

Girder A B

Siding attached Column


to columns
A B

FIGURE 32.1 Types of bracing: (a) relative; (b) nodal; (c) continuous; and (d) lean-on.

systems, such as diagonal bracing or shear walls, prevent the relative lateral movement of adjacent stories
or of adjacent points along the length of a member. Relative systems can be readily identified if a cut at
any location along the length of a braced member passes through the brace member itself. Nodal systems
control the movement only where they attach to the braced member and do not directly interact with
adjacent brace points. Cross-frames or diaphragms between two adjacent beams are considered nodal
braces. Continuous systems provide uninterrupted support along the entire length of a member, leaving
no unbraced length. Shear walls and roof or floor deck are examples of continuous bracing systems.
Lean-on systems rely on adjacent structural members to provide support. Lean-on bracing links together
adjacent structural members such that buckling of one member requires all members in the system to
buckle with the same lateral displacement.

32.3 Background
Structural bracing used to increase the strength of members must possess both sufficient strength and
stiffness [2]. Simple bracing design rules such as designing a brace to resist 2% of the member
compressive force address only the strength criterion. The stiffness of the brace along with the out-
of-straightness of the member has a direct effect on the magnitude of the brace force [1]. Design
recommendations based on perfectly straight members should not be used directly in design since
extremely large brace forces and displacements may result [3].

32.3.1 Member Out-of-Straightness and Brace Stiffness


Winter [1,2] developed the concept of a dual strength and stiffness criterion for the design of bracing used
to control instability. The required brace strength cannot be uniquely determined, but depends on both
the magnitude of the brace stiffness and member initial out-of-straightness. The relationship between
these parameters is illustrated for the relative column brace in Figure 32.2 and can be extended to other
types of bracing systems. In order to reach the Euler buckling load, Pe, the brace must possess a minimum
stiffness known as the ideal stiffness, bi. Figure 32.3 shows the relationship between the brace stiffness and
force. When the ideal stiffness is used (b ¼ bi), as the column load approaches Pe the sway deflections
become very large. Unfortunately, this results in very large brace forces since Pbr ¼ bD. At twice the ideal
stiffness (b ¼ 2bi), the brace force equals 0.4% of the column load when P ¼ Pe. For practical designs, the

Copyright 2005 by CRC Press


Structural Bracing 32-3

∆o P P
∆o ∆
 ∆

L
Initial out-
of-straightness

A
P P

FIGURE 32.2 Relative brace.

(a) 3i ∆ = ∆o (b)


1 1
2i 3i 2i
0.8 0.8
 = i ∆o P
Pbr 0.6 Pbr 0.6  = i

P P
0.4 L 0.4
∆T = ∆ + ∆o
∆o = 0.002L
0.2 0.2
P 0.4% P
0 0
0 4 8 12 16 20 0 0.5% 1.0% 1.5% 2.0%
∆T /∆o Pbr (% of P)

FIGURE 32.3 Effect of initial out-of-plumbness.

deflections and corresponding brace forces are kept small by using brace stiffnesses greater than the ideal
stiffness. The plots developed in Figure 32.3 were based on an assumed initial out-of-straightness equal to
0.002L. Larger out-of-straightness values linearly increase the magnitude of the brace forces.

32.3.2 Member Shortening


Compression elements such as columns or the compression flanges of beams shorten under externally
applied compressive forces. For relative tension braces, shortening increases the brace force requirements
by causing an apparent increase in the member out-of-straightness, as shown in Figure 32.4. As the
member being braced shortens, slack is introduced into the bracing. Lateral movement at the brace
points is necessary in order to return the tension brace to its original length prior to member shortening.
The increases in brace force due to shortening can be accounted for by adding the out-of-straightness
due to shortening to the initial out-of-straightness (see Example 32.1).

32.3.3 Member Inelasticity


The bracing requirements for relative braces are a function of the load on the member and the distance
between adjacent brace points and not the elasticity or inelasticity of the member. For a nodal bracing
system, there has been some debate about the effects of inelasticity on the bracing requirements.
Research, however, has indicated that inelasticity of the main members does not affect the bracing
requirements [3]. For continuous and lean-on bracing systems, the bracing requirements, which will be
presented later, are based on the elastic and inelastic stiffnesses of the braced members. For these bracing
systems, the influence of inelasticity on the buckling solution can be reasonably approximated using the
tangent modulus ET ¼ tE, where E is the elastic modulus and t is the inelastic stiffness reduction factor.

Copyright 2005 by CRC Press


32-4 Handbook of Structural Engineering

∆os = sway due to shortening = ∆sh tan 

∆sh = shortening

FIGURE 32.4 Effect of member shortening for relative bracing.

In the AISC load and resistance factor design (LRFD) specification, when the axial stress is less than one
third the yield stress, Fy, the column is classified as elastic (t ¼ 1.0). At greater stress levels, the stiffness
reduction factor is given by
   
Pu ðPu =Py Þ
t ¼ 7:38 log ð32:1Þ
Py 0:85
where Pu is the required column strength and Py is the column squash load.

32.3.4 Brace Connection Stiffness


The connections details used to attach structural bracing members can be of great importance when
designing or evaluating the overall performance of a bracing system. If the connections are flexible, the
stiffness of the overall bracing system can be significantly less than the stiffness of the bracing member
alone. The stiffness of a bracing system can be evaluated as springs in series using
1 1 X 1
¼ þ ð32:2Þ
bsys bbr bconn
The system stiffness, bsys, will always be less than the smaller of the brace member stiffness, bbr, and any
of the connection stiffnesses, bconn.

32.4 Safety Factors


The bracing design recommendations that will follow are based on ultimate strength. The loads on the
members being braces are assumed to be factored. Since both strength and stiffness are essential
requirements for adequate bracing, resistance factors for each are necessary. When using an LRFD
approach, the required stiffness is divided by a resistance factor of f ¼ 0.75 to obtain a conservative
requirement. The design brace force is based on factored loads and compared to the factored design
strength of the brace and its connections.

32.5 Column or Frame Bracing


Column or frame bracing systems can be relative, nodal, continuous, or lean-on. The design recom-
mendations for relative and nodal column bracing are based on an initial out-of-straightness
Do ¼ 0.002L, where L is the column length and a brace stiffness equal to twice the ideal stiffness. The
initial displacement, Do, is defined as the lateral offset between two adjacent brace points caused by
sources other than brace elongations from gravity loads or compressive forces. For example, Do may be
a displacement due to wind or other lateral forces, erection tolerances, or column shortening. If Do
differs from 0.002L, the brace force, Pbr, will change in direct proportion to the actual Do. In frame

Copyright 2005 by CRC Press


Structural Bracing 32-5

systems where a story may contain n0 columns, each having a random out-of-plumbness, an average
value for Do can be used [4]
pffiffiffiffiffi
Do ¼ 0:002L n0 ð32:3Þ

32.5.1 Column Buckling and Design Philosophy


For no-sway columns, the effective length factor, K, will be less than 1.0. Most designs conservatively use
K ¼ 1.0 and in most situations achieving K < 1.0 is not economical. For sway permitted columns, the
effective length K  1.0. Bracing used to prevent sway can reduce the effective length to K ¼ 1.0 and
achieve significant economic savings.
The bracing design criterion for columns is based on providing sufficient strength and stiffness to
allow a sway column to achieve the Euler buckling load corresponding to K ¼ 1.0. For columns that
possess nonzero end restraint, this does not correspond to the no-sway buckling load since for these cases
K is theoretically less than unity.
For flexural buckling modes, brace points attached in-line with the centroid of the structural member
are most effective. For torsional buckling modes, bracing must prevent twist of the cross-section to be
effective.

32.5.2 Relative Systems


AISC LRFD brace requirements for relative column bracing are
2Pu
bbr ¼ ð32:4Þ
fLb

Pbr ¼ 0:004Pu ð32:5Þ


where f ¼ 0.75, Pu is the required compressive strength of the column, and Lb is the required brace
spacing.
When the actual brace stiffness provided, bact, differs from the required value given in Equation 32.4,
the brace strength requirement can be modified using
X 1
Pbr ¼ 0:004 Pu ð32:6Þ
2  ðbbr =bact Þ

EXAMPLE 32.1

Relative Column Brace Design


A typical tension-only X-brace must stabilize three bents. Each bent carries a total factored load of
600 kip (125 þ 300 þ 175). Assume the floor acts as a rigid diaphragm and all Do ¼ 0.002L and all
columns are W14 53 (Ag ¼ 15.6 in.2).

125 k 300 k 175 k

12⬘


18⬘

Copyright 2005 by CRC Press


32-6 Handbook of Structural Engineering

X
Pu ¼ ð3 bentsÞð125 þ 300 þ 175Þ ¼ 1800 kip
Bracing shear force:
Pbr ¼ 0:004ð1800Þ ¼ 7:2 kip
Bracing shear stiffness:
2ð1800Þ
¼ 400 kip/ft
bbr ¼
0:75ð12Þ
Design recommendations assume brace shear force and stiffness are perpendicular to column. Therefore,
for an inclined threaded rod (A36 steel):
Strength:
7:2 kip
Pbr ¼ ¼ 8:64 kip ¼ 0:9ð36ÞAg ðAg Þreq d ¼ 0:27 in:2
cos y ’
Stiffness:
Ag E ðAg Þreq d ¼ 0:43 in:2 Controls
cos2 y ¼ 400 k/ft ’  
21:6 ft Use 34 in. dia. rod Ag ¼ 0:44 in:2
Consider effects of shortening

300 k
Dos ¼ sway due to shortening ¼ Dsh tan y
300ð12 12Þ
Dsh ¼ ¼ 0:095 in:
ð15:6Þð29,000Þ
 
.6⬘ 12
12⬘ 21 Dos ¼ 0:095 ¼ 0:063 in:
18
 Dos 0:063
¼ ¼ 0:00044
L ð12 12Þ
18⬘  
0:002 þ 0:0004
Pbr ¼ 8:64 kip ¼ 10:5 kip
0:002
ðAg Þreq d ¼ 0:32 in:2 < 0:43 in:2 stiffness still controls

32.5.3 Nodal Systems


Figure 32.5 shows the relationship between brace stiffness and buckling load for a column with three
equally spaced nodal braces. The exact solution taken from Timoshenko and Gere [5] illustrates the
increase in buckling load and changes in mode shapes as the brace stiffness increases.

Pcr
1.0

0.8 L

Pcr 0.6
Pe
Limit
0.4 full bracing
2EI
0.2 Pe =
L2 3.41
0
0 1 2 3 4
L/Pe

FIGURE 32.5 Discrete bracing.

Copyright 2005 by CRC Press


Structural Bracing 32-7

For nodal column bracing, the ideal brace stiffness is a function of the number of intermediate braces
[1,2]. For a single brace at midheight bi ¼ 2P/L. For many closely spaced braces the ideal stiffness
approaches bi ¼ 4P/L. Twice the ideal stiffness for the most severe case was adopted by the AISC LRFD
for many braces.
AISC LRFD brace requirements for nodal column bracing are
8Pu
bbr ¼ ð32:7Þ
fLb
Pbr ¼ 0:01Pu ð32:8Þ
where f ¼ 0.75, Pu is the required compressive strength of the column, and Lb is the required brace spacing.
For n equally spaced braces, the ideal stiffness can be approximated as
Pu
bi ¼ Ni ð32:9Þ
Lb
where Ni 4  2/n. Using the recommended stiffness equal to twice the ideal stiffness and applying the
resistance factor gives
2Pu
bbr ¼ Ni ð32:10Þ
fLb
Equation 32.10 is based on equally spaced braces and is unconservative for unequal spacings. The
required stiffness for unequal brace spacings can be obtained using Winter’s rigid bar model [6]. In this
model, the column is represented using rigid links with ficticious hinges at brace locations and the braces
are represented using simple springs. Under the applied load, displacements are imposed at brace
locations and equilibrium is enforced to obtain Ni. This technique is illustrated in Example 32.2. For
a single nodal brace at any location along the length of a column, with the longest segment defined as L
and the shorter segment as aL, Ni can be conservatively determined using
1
Ni ¼ 1 þ ð32:11Þ
a

EXAMPLE 32.2

Nodal Column Brace — Unequal Spacing


1. Introduce hinge at B and displace arbitrary distance D.
2. Sum forces and moments to obtain reactions at A and C.
3. Cut and sum moments at B on AB to obtain

P P
PD ¼ 0:6bDð0:4LÞ
A 0.6D
P
bi ¼ 4:16
0.4L L
D Ni ¼ 4:16
B
 D
Conservative approximation
0.6L
1
Ni ¼ 1 þ ¼ 2:5
C 0:4=0:6
0.4D

P P

The brace stiffness requirements for nodal bracing are inversely proportional to the unbraced
length, Lb. Closer-spaced braces require more stiffness because the derivations are based upon
allowing the column to reach a load that corresponds to buckling of the most critical unbraced length

Copyright 2005 by CRC Press


32-8 Handbook of Structural Engineering

with a K-factor equal to 1.0. In many instances, there are more potential brace points than necessary
to support the member forces required. Using the actual unbraced length may result in excessively
conservative stiffness requirements. Therefore, the maximum unbraced length that enables the col-
umn to reach the required loading, Lq, can be used. For example, say the column shown in
Figure 32.5 is supported against weak-axis buckling at three locations giving an unbraced length of
L. If a single brace at midheight giving an unbraced length of 1.5L would be sufficient to carry the
load on the column, then the required stiffness for the three braces could be conservatively estimated
using the permissible unbraced length of 1.5L in Equation 32.10 in place of the actual unbraced
length of L (see Example 32.3).

EXAMPLE 32.3

Nodal Column Brace Design


Two 8 ft cross-members brace a 30 ft WT5 19.5 compression member. Buckling about x–x axis con-
trols since brace flexural stiffness is much lower than axial stiffness. Fy = 36 ksi. Find the required brace
strength and stiffness.
Pu ¼ 120 k
2
n ¼ 2, Ni ¼ 4  ¼ 3
2
2ð200Þ F 48EI 10⬘ 4⬘
bbr ¼ 3 ¼ 13:33 k=in: ¼ ¼
0:75ð10 12Þ D ð8 12Þ3 4⬘
Pbr
10⬘
13:33ð96Þ3 4⬘ 4⬘
Ireq’d ¼ ¼ 8:5 in:4
48ð29,000Þ D 10⬘ Brace

Pbr ¼ 0:01ð200Þ ¼ 2 kip

2ð96Þ x x
Mbr ¼ ¼ 48 kip in:
4

32.5.4 Continuous Systems


Figure 32.6 shows the relationship between brace stiffness and buckling load for a continuously braced
column. The exact solution can be approximated using the following equation [3]:

qffiffiffiffiffiffiffi
2L Pe
Pcr ¼ Pe þ b ð32:12Þ
p

where Pe is the Euler buckling load, L is the column length, and b  is the brace stiffness per unit
length.
The continuous brace formulation given in Equation 32.12 can also be applied for equally spaced
, using
discrete braces by determining an equivalent brace stiffness per unit length, b

 ¼b n
b ð32:13Þ
L

where n is the number of braces within the column length, L. This method is accurate for two or more
discrete braces and is illustrated in Example 32.4.
Corrugated metal deck is a common type of continuous lateral bracing and acts like a shear
diaphragm with the properties of a relative brace. The stiffness and strength properties of the metal

Copyright 2005 by CRC Press


Structural Bracing 32-9

Pcr
20
Pe =  EI
2
EQ. 12.5
L2
L
15 n =3
 = k/ in. per in.
P EQ. 12.4
10
Pcr n =2

5
n =1
0
0 200 400 600 800
L2/Pe

FIGURE 32.6 Continuous bracing.

∆ Metal deck
F 2Pu
F br = =
∆  Lb
F = 0.004Pu
F/b
G⬘=
Lb ∆/Lb
∆/L
2Pu
G⬘req’d =
b

FIGURE 32.7 Continuous metal-deck bracing.

deck are generally defined in a per unit width basis (e.g., shear stiffness ¼ G 0 kip/rad per ft width).
The bracing requirements for the shear diaphragm can be determined from the relative brace
requirements presented in Section 32.5.2, as shown in Figure 32.7. Properties for corrugated deck can
be obtained from the Steel Deck Institute Diaphragm Design Manual [7]. The required shear dia-
phragm stiffness per unit width is
0 2Pu
Greq’d ¼ fb ð32:14Þ

Dividing the perpendicular brace force requirement by the diaphragm width gives the required shear
strength per unit width, Su
0:004Pu
Su ¼ ð32:15Þ
b
It should be noted that the brace force requirements given in Equation 32.13 are in addition to other
load demands placed on the diaphragm.

EXAMPLE 32.4

Nodal Column Braces as Effective Continuous Braces


Consider a column of length 3Lb with two equally spaced nodal braces giving an unbraced length of Lb.
Estimate the critical load of an ideally-braced column by approximating the nodal braces as an
equivalent continuous brace.

Copyright 2005 by CRC Press


32-10 Handbook of Structural Engineering

Ideal nodal brace stiffness required


  to buckle between brace points p2 EI
2 Pcr 3Pcr Pcr ¼
Ni 4  ¼3 bi ¼ Ni ¼ Lb2
n L Lb
Equivalent continuous stiffness using Equation 32.13
  
3Pcr 2 braces 2Pcr
b¼ ¼ 2 Lb
Lb 3Lb Lb
Critical load using Equation 32.12
s ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi L Lb
 ffi
Pcr 2ð3Lb Þ 2Pcr Pcr
Pcr ¼ þ 2 ¼ 1:011Pcr Lb
9 p Lb 9

32.5.5 Lean-On Systems


Lean-on bracing systems provide a means for some members in a structural system to rely on other
members for stability. One of the most commonly encountered examples of lean-on systems are frames
with ‘‘leaning’’ pin-ended columns connected to columns with nonzero end restraint. For these systems,
P
the P concept [8] can be used to design the members. This concept states that a system remains stable
P
so long as the sum of the applied loads, P, is less than the sum of the buckling strength of each
P
individual member, Pcr, and the load in any individual member is less than the load corresponding to
buckling between braces (no sway) for that member. This concept is illustrated in Example 32.5.

EXAMPLE 32.5

Lean-On Column Brace


Is the W10 33 capable of bracing the W12 58?
A36 steel, factored loads shown, fully braced out of plane
W10 33 W12 58
Ag ¼ 9.71 in.2 Ag ¼ 17.0 in.2 No sway capacity — From AISC Manual (KLy ¼ 8 ft)
Ix ¼ 171 in.4 Iy ¼ 107 in. W12 58: (fPn ¼ 482 k) > (Pu ¼ 450 k) OK

50 k 450 k P
Sway capacity — Using P concept

Column A (W12 58)

8 ft
Pu 450 1
¼ ¼ 0:735 > \ inelastic
Fy Ag ð36Þð17:0Þ 3
 
0:735
t ¼ 7:38ð0:735Þlog ¼ 0:342
W10 × 33

W12 × 58

0:85
8 ft
p2 ð29,000Þð107Þ
fPn ¼ 0:85ð0:342Þð0:877Þ ¼ 94 k
ð288Þ2
Column B (W10 33)
8 ft
Sway Pu 50 1
mode ¼ ¼ 0:143< \ elastic t ¼ 1:0
Fy Ag ð36Þð9:71Þ 3
p2 ð29,000Þð171Þ
fPn ¼ 0:85ð1:0Þð0:877Þ ¼ 440 k
B A ð288Þ2
P
Using P concept
X X
fPn ¼ 94 þ 440 ¼ 534 k > Pu ¼ 50 þ 450 ¼ 500 k OK

Copyright 2005 by CRC Press


Structural Bracing 32-11

32.5.6 Torsional Bracing


In order for a brace to effectively increase the load-carrying capacity of a member, it must restrain the
movement of the lowest buckling mode. Depending on the cross-section, the lowest buckling mode of
a compression member may be flexural (lateral), torsional, or a combined flexural–torsional. Bracing
against flexural modes must prevent the lateral translation of the member cross-section. Bracing against
torsional modes must restrain the twist of the cross-section. Bracing details such as a rod framing into
a wide-flange column web (Figure 32.8) resist lateral buckling about the weak axis but do not prevent
twist and are ineffective torsional braces.
For doubly symmetric sections such as wide-flange columns, weak-axis flexural buckling controls
for an unbraced member. Providing sufficient weak-axis lateral bracing, in some instances, will result
in the section being controlled by the torsional buckling mode. If bracing is provided that prevents
both translation and twist of a doubly symmetric cross-section, weak-axis buckling will always
control.
The torsional buckling load, PT, for a column restrained about an axis modified by t is given by the
following (see Figure 32.9) [5]:
Axis of restraint along weak axis of column:
 
tPey ðd 2 =4Þ þ ybr
2
þ GJ
PT ¼ 2 2 2
ð32:16Þ
ybr þ rx þ ry

Axis of restraint along strong axis of column:


 
tPey ðd 2 =4Þ þ ðIx =Iy Þxbr
2
þ GJ
PT ¼ 2 þ r2 þ r2 ð32:17Þ
xbr x y

where xbr, ybr are the coordinates of axis of restraint with respect to column centroid, d is the column
depth, and Pey is the Euler load based on a column length between points of zero twist.
To compensate for the assumption in the derivation of Equations 32.16 and 32.17 that the brace
is infinitely stiff, the maximum factored column load should be limited to 0.90PT [3].

FIGURE 32.8 Ineffective torsional column brace.

Copyright 2005 by CRC Press


32-12 Handbook of Structural Engineering

(a) (b)
Brace point (axis of restraint)
x x
ybr

d


FIGURE 32.9 Buckling of a column about a restrained axis: (a) lateral brace at flange and (b) buckled shape.

(a) (b)
Moment connection

Stiffener at least
half depth

FIGURE 32.10 Typical torsional brace details: (a) control twist with struts and (b) moment connection with stiffener.

If column loads greater than PT are required, torsional bracing must be provided. Two typical bracing
schemes are shown in Figure 32.10. For continuous girts with moment connections, twisting restraint is
provided. However, partial depth stiffeners should be used to control web distortion. The design
requirements for torsional bracing are based on the nodal requirements presented in Section 32.5.3
and are obtained by introducing equal and opposite brace forces on each flange. The magnitude of these
forces is based on the assumption that each flange carries one-half of the total column load. The resulting
brace moment, MT ¼ 0.5Pbrd. Using the angle of twist y ¼ D/d as shown in Figure 32.9b, the stiffness
requirement bT ¼ MT/y ¼ 0.5Pbrd2/D reduces to
bT ¼ 0:5bbr d 2 ð32:18Þ
where bbr is the nodal brace stiffness requirement from Section 32.5.3.

32.6 Beam Bracing


Structural bracing for beams can be generally classified as either a lateral or a torsional system. Lateral
beam bracing, like column bracing, can be relative, nodal, continuous, or lean-on while torsional systems
can be nodal or continuous. Cross-frames or diaphragms connecting adjacent girders control twist of the
cross-section and represent purely torsional systems. Similarly, joists attached to the compression flange
of a simply supported beam control twist by preventing the lateral movement of the compression flange
relative to the tension flange and represent a purely lateral system. Some bracing systems, such as
a composite slab attached to the top flanges of a beam using shear connectors, simultaneously provide
both lateral and torsional restraints. Combined lateral and torsional systems have been shown to be more
effective braces than either lateral or torsional systems alone for beams with uniform moment [9,10].
In order to be an effective brace, both lateral and torsional systems must prevent the relative dis-
placement of the top and bottom of a beam (i.e., twist of the cross-section). Properly designed cross-
frames interconnecting adjacent beams are considered brace points when evaluating the buckling
strength of the beams. Even though the beams can move laterally, the cross frames are still effective
braces because they prevent twist of the cross-section. Both tests and theory have confirmed this
fact [11,12].

Copyright 2005 by CRC Press


Structural Bracing 32-13

32.6.1 Inflection Points as Brace Points


Inflection points are sometimes mistakenly identified as brace points in restrained beams. This often
occurs when the top flange of a beam is laterally braced by a slab or joists all along the span while
the bottom flange remains unbraced. Even in these instances, the inflection point cannot be consi-
dered a brace point. The justification for this can be illustrated by comparing the two beams shown in
Figure 32.11. Beam A has a moment on one end with an unbraced length Lb ¼ L while Beam B has equal
and opposite end moments resulting in an inflection point at midspan and an unbraced length Lb ¼ 2L.
The buckling moment for Beam B is 68% of Beam A. If the inflection point were a brace point, the
critical moment for both beams would be identical. A plan view of the buckled shape of Beam B
illustrates how the top and bottom flanges move in opposite directions at midspan. Even an actual brace
on one flange at the inflection point does not provide effective bracing at midspan because twist is not
restrained [12].

32.6.2 Lateral Bracing


The primary factors influencing the effectiveness of lateral beam bracing are
 Number or spacing of braces
 Vertical position of the braces on the cross-section
 Vertical position of the loads on the cross-section
 Moment gradient on the beam
Increasing the number of braces along the length of a beam can increase the load-carrying capacity by
reducing the unbraced length. Like column bracing, however, the brace stiffness requirements increase
with the number of braces. This is because the design recommendations assume the beam must reach
a load corresponding to buckling between brace points. Similarly, if more brace points are provided than
necessary to support the required loads, the maximum permissible unbraced length, Lq, may be used in
place of the actual unbraced length, Lb.
The vertical position of a lateral brace along the height of a beam significantly affects the effective-
ness of lateral bracing. Lateral bracing is most effective when positioned at the compression flange of

100

Brace
Beam A
L
68

L
Beam B
L

68
Moment diagrams

Mid-depth
Top flange

Bottom flange
Ineffective
Plan of buckled shape of Beam B midspan brace

FIGURE 32.11 Beam with inflection point.

Copyright 2005 by CRC Press


32-14 Handbook of Structural Engineering

(a) Pcr (b) Pcr

Compression
flange
Compression
flange

FIGURE 32.12 Lateral buckling of cantilever and simple beams: (a) max moment at max twist and (b) zero
moment at max twist.

FIGURE 32.13 Effect of load and brace positions.

a beam. In most configurations, this places the brace the furthest distance from the center of twist of the
buckled cross-section as shown in Figure 32.12a. The exception to this rule is for cantilever beams
(Figure 32.12b) where the position of the buckled cross-section reveals that a brace located at the tension
flange is most beneficial.
When a beam is bent in reverse curvature, the compression and tension flanges switch at the point of
inflection. Beams such as these, in which both the top and bottom flanges encounter compression along the
span, have more severe bracing requirements than beams where compression resides only in one flange [3].
In these situations, lateral bracing is required on both flanges to prevent twist of the cross-section.
The vertical position of transverse loads on beams has significant influence on the effectiveness of
lateral bracing. Loads applied higher on the cross-section, such as at the top flange, have more pro-
nounced destabilizing effects while loads applied lower on the cross-section tend to provide added
stability when compared to centroidal loading. The effect of top-flange loading is even greater when
lateral bracing is located near the centroid of the section. In these situations, the center of twist shifts to
a position closer to mid-depth and the centroidal brace becomes almost totally ineffective as shown in
Figure 32.13. Therefore, centroidal lateral beam braces are not recommended due to the effects of both
cross-section distortion and load position.
The load position effect described above is based on the assumption that the load remains vertical and
passes through the original point of contact on the member as it buckles. For many structural systems,

Copyright 2005 by CRC Press


Structural Bracing 32-15

FIGURE 32.14 Tipping effect and cross-section distortion.

the load transferred to beams is applied through secondary members or a floor slab. When loading is
through a slab, for example, a restoring torque is created by a tipping effect during buckling as illustrated
in Figure 32.14. This tipping effect has been shown to significantly increase the lateral buckling capacity
even if the slab is only resting (not positively attached) on the top flange [12]. Unfortunately, the benefits
of tipping are severly limited by distortion of the cross-section and are difficult to quantify. As a result,
the beneficial effects of tipping are generally neglected.
The AISC LRFD lateral beam brace requirements were based on the following design recommenda-
tions developed by Yura [3]:
The brace stiffness requirement for both relative and nodal beam bracings is
2Ni ðCb Pf ÞCt Cd
bbr ¼ ð32:19Þ
fLb
The brace strength requirement for relative bracing is
Mu Ct Cd
Pbr ¼ 0:004 ð32:20Þ
hoh
and for nodal bracing it is
Mu Ct Cd
Pbr ¼ 0:01 ð32:21Þ
hoh
where
Ni ¼ 1.0 for relative bracing
¼ (4  2/n) for nodal bracing
n ¼ number of intermediate braces
Pf ¼ beam compressive flange force
¼ p2 EIyc =Lb2
Iyc ¼ out-of-plane moment of inertia of the compression flange
Ct ¼ top-flange loading factor
¼ 1.0 for centroidal loading
¼ 1 þ (1.2/n) for top-flange loading
Cd ¼ 1 þ (MS/ML)2 for reverse-curvature bending
¼ 1.0 for single-curvature bending
MS ¼ smallest moment causing compression in each flange
ML ¼ largest moment causing compression in each flange
Cb ¼ nonuniform moment modification factor
12:5Mmax
¼
2:5Mmax þ 3MA þ 4MB þ 3MC
Mmax ¼ absolute value of maximum moment in unbraced segment
MA ¼ absolute value of moment at quarter point of unbraced segment
MB ¼ absolute value of moment at midspan of unbraced segment
MC ¼ absolute value of moment at three-quarter point of unbraced segment

Copyright 2005 by CRC Press


32-16 Handbook of Structural Engineering

The brace force requirements were developed assuming an initial lateral displacement of the
compression flange equal to 0.002Lb and vary in direct proportion to the actual out-of-straightness.
The term 2NiCt can be conservatively approximated as 10 for any number of nodal braces and 4 for
any number of relative braces. The term CbPf can also be conservatively approximated as Mu/hoh. Using
the worst-case top-flange loading (Ct ¼ 2.0) and the previous assumptions yields the AISC LRFD brace
requirements for lateral beam bracing (Example 32.6):
AISC LRFD brace requirements for relative lateral beam bracing:
4Mu Cd
bbr ¼ ð32:22Þ
fLb hoh

Mu Cd
Pbr ¼ 0:008 ð32:23Þ
hoh
AISC LRFD brace requirements for nodal lateral beam bracing:
10Mu Cd
bbr ¼ ð32:24Þ
fLb hoh

Mu Cd
Pbr ¼ 0:02 ð32:25Þ
hoh
where
f ¼ 0.75
Mu ¼ required flexural strength
ho ¼ distance between flange centroids
Cd ¼ 1 þ (MS/ML)2 for reverse-curvature bending
¼ 1.0 for single-curvature bending
Lb ¼ distance between braces

EXAMPLE 32.6

Relative Lateral Beam Brace


Design the diagonals of the top flange horizontal truss to stabilize the five simply-supported 90 ft girders.
The factored moment at midspan Mu ¼ 1200 k ft and Fy ¼ 36 ksi for bracing.
Top flange
4 × 9 ft = 36 ft

Section view

5 × 18 ft = 90 ft 1 × 10
Plan view
Stiffness
1
2 × 48

4:0ð1200 12Þð1:0Þ 1 × 16
bbr ¼ ð2:5 girdersÞ ¼ 18:1 k=in:
0:75ð18 12Þð49Þ ho = 49 in.
    2
AE 2 Abr ð29,000Þ 1
cos y ¼ p ffiffi
ffi pffiffiffi ¼ 18:1
L br 9 12 5 5

Abr ¼ 0:75 in:2 Controls

Copyright 2005 by CRC Press


Structural Bracing 32-17

Strength
ð1200 12Þð1:0Þ
Pbr ¼ 0:008 ð2:5 girdersÞ ¼ 5:88 k
49
pffiffiffi
5:88 5 1
Abr ¼ ¼ 0:41 in:2 Use L2 2 ðAg ¼ 0:944 in.2 Þ
ð0:9 36Þ 4

32.6.3 Torsional Bracing


The primary factors influencing the effectiveness of lateral beam bracing have relatively little influence on
the design of torsional beam bracing. Unlike lateral beam bracing, the number of braces, brace location
on the cross-section, and load position are relatively unimportant when sizing a torsional brace. Not only
is a torsional brace equally effective if attached to the top or bottom flange, but beams in reverse
curvature do not alter the torsional brace requirements. The effectiveness of a torsional brace, however, is
greatly affected by distortion of the cross-section as illustrated in Figure 32.15. Although the torsional
brace prevents twist at the top flange, distortion of the web permits a relative displacement between the
two flanges. A web stiffener located at the brace location is often used to control the distortion.
AISC LRFD brace requirements for torsional beam bracing:
bT
bTb ¼ ð32:26Þ
ð1  ðbT =bsec ÞÞ

0:024Mu L
Mbr ¼ ð32:27Þ
nCb Lb
where
 
3:3E 1:5hoh tw3 ts bs3
bsec ¼ þ ð32:28Þ
hoh 12 12

2:4LMu2
bT ¼ ð32:29Þ
fnEIy Cb2
where

f ¼ 0.75
L ¼ span length
n ¼ number of nodal braced points within span
E ¼ modulus of elasticity
Iy ¼ out-of-plane moment of inertia of beam
Cb ¼ nonuniform moment modification factor
tw ¼ thickness of beam web

Torsional brace

Web

FIGURE 32.15 Cross-section distortion.

Copyright 2005 by CRC Press


32-18 Handbook of Structural Engineering

Stiffener extends full


height of brace
tw

4tw

Attach when brace is


attached to flange

FIGURE 32.16 Web stiffener details for torsional beam bracing.

ts ¼ thickness of web stiffener(s)


bs ¼ stiffener width for one-sided stiffeners, which is twice the individual stiffener width for pairs of
stiffeners
bT represents the torsional stiffness of the brace member itself. For flexible connections, Equation 32.2
should be used. bsec is the torsional stiffness associated with the beam web and any transverse web stiffeners
that may be present (web distortional stiffness). The required torsional brace stiffness, bTb, will be negative
if bsec < bT, which indicates the brace will not be effective due to insufficient web distortional stiffness. The
coefficient 2.4 in Equation 32.29 comes from using twice the ideal stiffness with a 20% increase to account
for top-flange loading. The required strength, Mbr, assumes an initial twist of 1 (0.0175 rad).
When web stiffeners are required, the AISC LRFD specification requires they extend the full depth of
the brace as shown in Figure 32.16. When the brace is attached to the beam flange the stiffeners must also
be attached to the flange. When the brace is not attached to the beam flange, the stiffener may be
terminated at a distance equal to 4tw from the flange.
The continuous torsional beam bracing requirements use the same formulations as the nodal bracing
requirements. For continuous bracing, the term L/n in Equations 32.27 and 32.29 is taken equal to 1.0
and the unbraced length, Lb, in Equation 32.27 is taken equal to Lq, the maximum unbraced beam length
necessary to carry the required factored loads.
For singly symmetric cross-sections, an effective moment of inertia, Ieff, is used in place of Iy as given by
yt
Ieff ¼ Iyc þ Iyt ð32:30Þ
yc
where Iyc is the compression flange out-of-plane moment of inertia, Iyt is the tension flange out-of-plane
moment of inertia, yc is the distance from centroid to extreme compression fiber, and yt is the distance
from centroid to extreme tension fiber.
The torsional stiffness of typical beam-type braces is shown in Figure 32.17. Adjacent girders connected on
the top flanges with decking, for example, will buckle in the same direction and develop the reverse-curvature
stiffness of the brace. The stiffnesses of typical truss-type diaphragm systems are shown in Figure 32.18
and are based on elastic truss analyses. For X-systems designed for tension only, the horizontal members
in the truss are not required. For K-brace systems, only one horizontal member is necessary.

32.7 Faulty Details


Numerous structural failures have occurred because of the structural arrangement shown in Figure 32.19.
The beam (or truss) is continuous over the top of the column. The critical components are: column in
compression, compression in the bottom flange of the beam or chord of the truss, and no bottom flange
bracing at point a and possibly point b. The sway at the top of the column shown in Sect B–B can result
in a K-factor much greater than 2.0. The bottom flange of the beam can possibly provide bracing to the

Copyright 2005 by CRC Press


Structural Bracing 32-19

Ib

 Mbr 

6EIb 2EIb
T = T =
S S

FIGURE 32.17 Stiffness formulas for diaphragms.

S
P P
/S 2ES2hb2r
Tension PL
+2 T =
system hbr 2Ld3 S3
+
P P Ad Ah
–P
2Phbr 2Phbr
S S

S
P P
/S
Compression PL
system +2 AdES2 hb2r
hbr T =
–2
PL L3d
P /S P

2Phbr 2Phbr
S S
S
P P
2ES2hb2r
T =
8L3d S3
–2

/S

K-brace hbr
+
PL
PL

Ad Ah
+2
/S

P P
+P –P
2Phbr 2Phbr
S S

Ad = Area of diagonal members
∆ + ∆b
Ah = Area of horizontal members  T = M =
Ld = Length of diagonal members  S

∆b

FIGURE 32.18 Stiffness formulas for cross-frames.

Copyright 2005 by CRC Press


32-20 Handbook of Structural Engineering

V V
b Joist
a b

Compression B
flange

Sect B–B
P

FIGURE 32.19 Structural detail with probable instability.

P
Siding
a

LG

Girt LG LC

LG

In-plane Out-of-plane

FIGURE 32.20 Inadequate torsional bracing detail.

top of the column if there are braces at point b and consideration is given to the compression in the
flange when evaluating its stiffness. In general, a brace, such as a bottom chord extension from the joist,
should be used at point a. Beam web stiffeners at the column location will also be effective unless bottom
flange lateral buckling is critical.
Another common faulty bracing detail is shown in Figure 32.20. The girts that frame into the column
flange prevent weak-axis translation at the braced flange. Since the girts are discontinuous, they will not
prevent twist of the cross-section and will force the column to buckle about a restrained axis (see also
Figure 32.9b). For this column, there are three possible buckling modes: strong-axis flexural buckling
(Lb ¼ KLC), weak-axis flexural buckling (Lb ¼ KLG), and torsional buckling about a restrained axis
(Lb ¼ KLC, ybr ¼ a assuming no twist at column ends).

Copyright 2005 by CRC Press


Structural Bracing 32-21

Nomenclature
A cross-sectional area of primary MS smallest moment causing compression
member in each flange along beam length
Abr cross-sectional area of brace member MT torsional brace moment
Ad area of diagonal member in Mu required bending strength
cross-frame Ni brace stiffness coefficient for nodal
Ah area of horizontal member in braces
cross-frame Pbr brace force
Ag gross cross-sectional area Pcr member buckling load
Cb bending coefficient dependent on Pe Euler column buckling load
moment gradient Pey Euler buckling load based on distance
Cd reverse-curvature bending factor between points of zero twist
Ct top-flange loading factor Pf beam compressive flange force
E modulus of elasticity PT torsional buckling load
ET tangent modulus of elasticity Pu required compressive strength of
Fy specified minimum yield stress column
G shear modulus of elasticity Py column squash load
G0 diaphragm shear stiffness per unit S cross-frame or diaphragm length
width Sx strong-axis section modulus
I moment of inertia Su required diaphragm shear strength
Ieff effective moment of inertia for singly per unit width
symmetric beam sections V shear force
Ireq’d required moment of inertia b orthogonal distance between point of
Ix, Iy moment of inertia about strong and restraint and weak axis of member
weak axes, respectively bs stiffener width for one-sided stiffeners
Iyc, Iyt out-of-plane moment of inertia of d member depth
compression and tension flanges, fb bending stress
respectively hbr deight of cross frame
J torsion constant for section torsional brace
K column effective length factor hoh distance between flange centroids
L member length n rumber of braces within span
Lb required brace spacing or laterally n0 rumber of columns in a story
unbraced length; length between rx, ry radius of gyration about strong and
points that are either braced against weak axes, respectively
lateral displacement of the compres- tw thickness of beam web
sion flange or braced against twist of ts thickness of web stiffener
the cross-section xbr, ybr coordinates of axis of restraint with
Lq maximum unbraced length for respect to column centroid
required forces yc, yt coordinates with respect to centroid of
MA absolute value of moment at quarter expreme compression and tension
point of unbraced beam segment fibers, respectively
MB absolute value of moment at midspan b continuous brace stiffness per unit
of unbraced beam segment length
MC absolute value of moment at bact stiffness provided by brace member
three-quarter point of unbraced bbr required lateral brace stiffness
beam segment bconn stiffness of brace connection
ML largest moment causing compression bsec web distortional stiffness including
in each flange along beam length any transverse web stiffeners if present
Mmax absolute value of max moment in bsys stiffness of brace system
unbraced beam segment bT nodal torsional brace stiffness

Copyright 2005 by CRC Press


32-22 Handbook of Structural Engineering

bTb required nodal torsional brace stiffness DT total column sway deflection
including web distortion f resistance factor
D translational displacement y complementary angle between
Do column initial out-of-straightness diagonal brace and axial member or
Dos column out-of-straightness due to twist of member cross-section
shortening t inelastic stiffness reduction
Dsh shortening or compression element factor

References
[1] Winter, G. (1958), ‘‘Lateral Bracing of Columns and Beams,’’ Trans. ASCE, Vol. 125, Part 1,
pp. 809–825.
[2] Winter, G. (1960), ‘‘Lateral Bracing of Columns and Beams,’’ Proc. ASCE, Vol. 84 (ST2), pp. 1561-
1–1561-22.
[3] Yura, J.A., Bracing, in Stability Design Criteria for Metal Structures, 5th Edition, Galambos, T.V.
Ed.; John Wiley & Sons, Inc., New York, 1998; Chapter 12.
[4] Chen, S. and Tong, G. (1994), ‘‘Design for Stability: Correct Use of Braces,’’ Steel Struct.,
J. Singapore Struct. Steel Soc., Vol. 5, No. 1, pp. 15–23.
[5] Timoshenko, S. and Gere, J. (1961), Theory of Elastic Stability, McGraw-Hill Book Company,
New York.
[6] Yura, J.A. (1994), ‘‘Winters Bracing Model Revisited,’’ 50th Anniversary Proc., Struc. Stability
Research Council, pp. 375–382.
[7] Luttrell, L.D. (1987), Diaphragm Design Manual, 2nd Edition, Steel Deck Institute, Fox River
Grove, IL.
[8] Yura, J.A. (1971), ‘‘The Effective Length of Columns in Unbraced Frames,’’ Eng. J. Am. Inst. Steel
Const., Vol. 8, No. 2, pp. 37–42.
[9] Mutton, B.R. and Trahair, N.S. (1973), ‘‘Stiffness Requirements for Lateral Bracing,’’ ASCE
J. Struct. Div., Vol. 99, No. ST10, pp. 2167–2182.
[10] Tong, G. and Chen, S. (1988), ‘‘Buckling of Laterally and Torsionally Braced Beams,’’ J. Const. Steel
Res., Vol. 11, pp. 41–55.
[11] Flint, A.R. (1951), ‘‘The Stability of Beams Loaded Through Secondary Members,’’ Civil Eng.
Public Works Rev., Vol. 46, No. 537–8 (see also pp. 259–260).
[12] Yura, J.A. (1993), ‘‘Fundamentals of Beam Bracing,’’ Proc. Struc. Stability Research Council Annual
Technical Session, ‘‘Is Your Structure Suitably Braced?’’ Milwaukee, April, 20 pp.

Further Reading
Akay, H.U., Johnson, C.P. and Will, K.M. (1977), ‘‘Lateral and Local Buckling of Beams and Frames,’’
ASCE J. Struct. Div., Vol. 103, No. ST9, pp. 1821–1832.
Ales, J.M. and Yura, J.A. (1993), ‘‘Bracing Design for Inelastic Structures,’’ Proc., SSRC Conf. ‘‘Is Your
Structure Suitably Braced?,’’ Milwaukee, April, pp. 29–37.
American Institute of Steel Construction (1995), Manual of Steel Construction: Load & Resistance Factor
Design, Vol. 1, 2nd Edition, Chicago.
American Society of Civil Engineers (1971), ‘‘Commentary on Plastic Design in Steel,’’ ASCE Manual
No. 41, 2nd Edition, New York.
Essa, H.S. and Kennedy, D.J.L. (1995), ‘‘Design of Steel Beams in Cantilever-Suspended-Span
Construction,’’ ASCE J. Struct. Div., Vol. 121, No. 11, pp. 1667–1673.
Gil, H. (1966), ‘‘Bracing Requirements for Inelastic Steel Members,’’ PhD dissertation, The University of
Texas at Austin, May, 156 pp.
Helwig, T.A., Yura, J.A., and Frank, K.H. (1993), ‘‘Bracing Forces in Diaphragms and Cross Frames,’’
Proc., SSRC Conf., ‘‘Is Your Structure Suitably Braced?’’ Milwaukee, April, pp. 129–140.

Copyright 2005 by CRC Press


Structural Bracing 32-23

Horne, M.R. and Ajmani, J.L. (1971), ‘‘Design of Columns Restrained by Side Rails,’’ Struct. Eng., Vol. 49,
No. 8, pp. 329–345.
Horne, M.R. and Ajmani, J.L. (1972), ‘‘Failure of Columns Laterally Supported on One Flange,’’ Struct.
Eng., Vol. 50, No. 9, pp. 355–366.
Lutz, A.L. and Fisher, J. (1985), ‘‘A Unified Approach for Stability Bracing Requirements,’’ Eng. J., Am.
Inst. Steel Constr., Vol. 22, No. 4, pp. 163–167.
Medland, I.C. and Segedin, C.M. (1979), ‘‘Brace Forces in Interbraced Column Structures,’’ ASCE
J. Struct. Div., Vol. 105, No. ST7, pp. 1543–1556.
Milner, H.R. and Rao, S.N. (1978), ‘‘Strength and Stiffness of Moment Resisting Beam-Purlin
Connections,’’ Civil Eng. Trans., Inst. of Engrg, Australia, CE 20(1), pp. 37–42.
Nakamura, T. (1988), ‘‘Strength and Deformability of H-Shaped Steel Beams and Lateral Bracing
Requirements,’’ J. Const. Steel Res., Vol. 9, 217–228.
Plaut, R.H. (1993), ‘‘Requirements for Lateral Bracing of Columns with Two Spans,’’ ASCE J. Struct. Div.,
Vol. 119, No. 10, pp. 2913–2931.
Pincus, G. (1964), ‘‘On the Lateral Support of Inelastic Columns,’’ Eng. J., AISC, Vol. 1, No. 4,
pp. 113–115.
Salmon, C.S. and Johnson, J.E. (1996), Steel Structures — Design and Behavior, 4th Edition, Harper and
Row, New York.
Taylor, A.C. and Ojalvo, M. (1966), ‘‘Torsional Restraint of Lateral Buckling,’’ ASCE J. Struct. Div.,
Vol. 92, No. ST2, pp. 115–129.
Timoshenko, S. and Gere, J. (1961), Theory of Elastic Stability, McGraw-Hill Book Company, New York.
Tong, G. and Chen, S. (1987), ‘‘Design Forces of Horizontal Inter-Column Braces,’’ J. Const. Steel Res.,
Vol. 7, pp. 363–370.
Tong, G. and Chen, S. (1989), ‘‘The Elastic Buckling of Interbraced Girders,’’ J. Const. Steel Res., Vol. 14,
pp. 87–105.
Trahair, N.S. and Nethercot, D.A. (1984), ‘‘Bracing Requirements in Thin-Walled Structures,’’ Chapter 3,
Developments in Thin-Walled Structures, Vol. 2, J. Rhodes and A.C. Walker, Eds., Elsevier,
Amsterdam, pp. 93–130.
Wang, Y.C. and Nethercot, D.A. (1989), ‘‘Ultimate Strength Analysis of Three-Dimensional Braced
I-Beams,’’ Proc. Inst. of Civil Engrg, Part 2, 87, March, London, pp. 87–112.
Winter, G. (1960), ‘‘Lateral Bracing of Columns and Beams,’’ Trans. ASCE, Vol. 125, Part 1, pp. 809–825.
Yura, J.A. (1995), ‘‘Bracing for Stability-State-of-the-Art,’’ Proceedings, Structures Congress XIII, ASCE,
Boston, April, pp. 88–103.
Yura, J.A. and Phillips, B.A. (1992), ‘‘Bracing Requirements for Elastic Steel Beams,’’ Research Report
1239-1, Center for Transportation Research, Univ. of Texas at Austin, May, 73 pp.
Yura, J.A., Phillips, B., Raju, S., and Webb, S. (1992), ‘‘Bracing of Steel Beams in Bridges,’’ Research
Report 1239-4F, Center for Transportation Research, Univ. of Texas at Austin, October, 80 pp.
Zuk, W. (1956), ‘‘Lateral Bracing Forces on Beams and Columns,’’ ASCE J. Eng. Mech. Div., Vol. 82,
No. EM3, pp. 1032-1–1032-11.

Copyright 2005 by CRC Press


Copyright 2005 by CRC Press

You might also like