Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Compression Flange

Related terms:

Lateral-Torsional Buckling, Girder, Flanges, Bending Moment, Fiber-Reinforced


Polymer, Buckling Mode, Steel Beam

View all Topics

Buckling/Plastic Collapse Behavior and


Strength of Plate Girders Subjected to
Combined Bending and Shear Loads
Tetsuya Yao, Masahiko Fujikubo, in Buckling and Ultimate Strength of Ship and
Ship-Like Floating Structures, 2016

7.2.4.2 Collapse under pure bending


Under pure bending, the compression flange of the plate girder may buckle tor-
sionally. On the other hand, the flange of the DB model in the compression side of
bending is subjected to thrust and may locally buckle as a continuous panel under
thrust. The difference in deformation of these two flanges are schematically shown
in Fig. 7.28.

Fig. 7.28. Local buckling of wide flange [11].

Two cases are analyzed here: the DB model and the same model but with free edge
flange of the same size. Deformation and spread of yielding of former DB model in
pure bending are shown in Fig. 7.29, and the moment-rotation angle relationships
for two cases in Fig. 7.30.
Fig. 7.29. Deformation and spread of yielding (DB model in bending) [9]. (A) Ultimate
strength (M/MP = 0.67). (B) Postultimate strength range (M/MP = 0.51).

Fig. 7.30. Bending moment-rotation angle relationships (DB model in bending) [9].

The change in bending stress distribution is also shown in Fig. 7.29. It is known that
compressive stress in the buckled part does not increase and the neutral axis moves
toward the tension side of bending.

Significant difference is observed in the bending moment-rotation angle relation-


ship depending on the condition of the compression flange. In the case of a continu-
ous flange, local panel buckling takes place and the compressive ultimate strength is
attained in the flange. The continuous plating condition evidently ensures the higher
ultimate strength because the rotation of the compression flange is constrained
owing to the continuity condition of the flange. On the other hand, decrease in
the capacity beyond the ultimate strength is much rapid in the case of a continuous
flange. This may be because of the concentration of buckling deformation at a certain
cross-section of the compression flange and the resulting elastic unloading in the
remaining part of the span. However, after a while, moment-rotation curves for the
two cases become almost the same.

> Read full chapter

Design Examples of Steel and


Steel-Concrete Composite Bridges
Ehab Ellobody, in Finite Element Analysis and Design of Steel and Steel-Concrete
Composite Bridges, 2014

4.5.6 Check of Lateral Torsional Buckling of the Plate Girder


Compression Flange
To check the safety of the upper compression flange against lateral torsional buck-
ling, we have to calculate the elastic critical moment for lateral torsional buckling
(Mcr). However, to calculate Mcr, we have to evaluate the effective buckling length
(unsupported length) of the compression flange of this investigated pony bridge (le),
which can be calculated as follows:

, where Ic is the inertial of the compression flange (see Figure 4.147) about z-z axis, a
is the spacing between cross girders, and is the flexibility of the U-frame reasonably
assumed = 0.00006 mm/N for pony bridges.
Figure 4.147. Check of lateral torsional buckling of plate girder.

(should be not less than a = 450 cm)

Considering the cross section at midspan shown in Figure 4.147, we can calculate
Mcr as follows:

We can now check the safety against lateral torsional buckling following the rules
specified in EC3 [1.27, 2.11] as follows:

Given: MEd = 20, 285.7 kN m, Wy = 96, 417.7 cm3

> Read full chapter


Applied Loads and Stability of Steel and
Steel-Concrete Composite Bridges
Ehab Ellobody, in Finite Element Analysis and Design of Steel and Steel-Concrete
Composite Bridges, 2014

3.8.3 Lateral Torsional Buckling of Plate Girders in Bending


When beams and plate girders are subjected to bending moment, the compression
flange will be subjected to lateral torsional buckling. The lateral torsional buckling
of the compression flange depends on the loading conditions, lateral restraint
conditions, and geometries of the compression flange. EC3 [1.27,2.11] recommends
that a laterally unrestrained member subject to major axis bending should be verified
against lateral torsional buckling as follows:

(3.38)

where MEd is the design value of the moment and Mb,Rd is the design buckling
resistance moment. Beams with sufficient restraint to the compression flange are
not susceptible to lateral torsional buckling. In addition, beams with certain types of
cross sections, such as square or circular hollow sections, fabricated circular tubes,
or square box sections, are not susceptible to lateral torsional buckling. The design
buckling resistance moment of a laterally unrestrained beam should be taken as

(3.39)

where Wy is the appropriate section modulus, which is taken as Wpl,y for class 1 or 2
cross sections or Wel,y for class 3 cross sections or Weff,y for class 4 cross sections,
and LT is the reduction factor for lateral torsional buckling. It should be noted
that, according to EC3 [1.27,2.11], determining Wy holes for fasteners at the beam
end need not be taken into account. Also, for bending members of constant cross
section, the value of LT for the appropriate nondimensional slenderness should be
determined from

(3.40)

(3.41)

(3.42)

where LT is an imperfection factor and Mcr is the elastic critical moment for lateral
torsional buckling. Mcr is based on gross cross-sectional properties and takes into
account the loading conditions, the real moment distribution, and the lateral re-
straints. The imperfection factor LT corresponding to the appropriate buckling curve
can be taken from Table 3.18 as specified in EC3 [2.11]. The recommendations for
buckling curves are given in Table 3.19 as specified in EC3 [2.11].

Table 3.18. Recommended Values for Imperfection Factors for Lateral Torsional
Buckling Curves as Given by EC3 [3.5]

Buckling curve a b c d
Imperfection factor 0.21 0.34 0.49 0.76
LT

Table 3.19. Values for Lateral Torsional Buckling Curves for Different Cross Sections
Recommended by EC3 [3.5]

Cross Section Limits Buckling Curve


Rolled I-sections h/b ≤ 2h/b > 2 ab
Welded I-sections h/b ≤ 2h/b > 2 cd
Other cross sections – d

> Read full chapter

Cracked structures and residual


strength
F WANG, W C CUI, in Condition Assessment of Aged Structures, 2008

8.8.3 Ultimate strength of ship hulls


Vasta (1958) assumed that the ship hull would reach the ultimate limit state when
the compression flange, i.e. the upper deck in the sagging condition or the bottom
plating in the hogging condition, collapses, and that the relationship between the
bending moment and curvature is linear. Caldwell (1965) took into account buckling
in compression and yielding in tension. The ship hull cross-section was idealized as
an equivalent section with uniform plate thickness in deck, bottom or sides. When
the ship hull reached the ultimate limit state, the entire material in compression was
assumed to have reached its ultimate buckling strength and the entire material in
compression was assumed to have reached full yielding. However, in the immediate
vicinity of the final neutral axis, the side shells will often remain in the elastic
state up to the overall collapse of the hull girder. So Paik and Mansour (1995)
developed Caldwell's method further. They assumed a more credible distribution
of longitudinal stresses of the hull cross-section in the collapse state: see Fig. 8.35.
Based on this assumption, the formulae for prediction of the ultimate strength can
be derived.

8.35. Assumed distribution of longitudinal stresses in a hull cross-section at the


overall collapse state: (a) sagging, (b) hogging.

In Fig. 8.35 and the discussion that follows, AB is the total sectional area of the outer
bottom, A'B is the total sectional area of the inner bottom, AD is the total sectional
area of the deck, AS is the half-sectional area of all sides, D is the hull depth, DB is the
height of the double bottom, g is the neutral axis position above the base line in the
sagging condition or below the deck in the hogging condition, H is the depth of the
hull section in the linear elastic state, Muh and Mus are the ultimate bending moments
in hogging or sagging conditions respectively, yB, yB, yD and ys are the yield
strength of the outer bottom, inner bottom, deck and side shell, respectively, and
uB, u , uD and us are the ultimate buckling strength of the outer bottom, inner

bottom, deck and side shell, respectively.

If the x–y coordinates are taken as shown in Fig. 8.35, the stress distribution can be
expressed as follows:

In sagging condition:

8.30

In hogging condition:

8.31

From the condition that no axial force acts on the hull girder, the depth of the
collapsed sides (D – H) can be obtained from

8.32

8.33

where

The position of the neutral axis where the longitudinal stress is zero can be deter-
mined by substituting Eqs 8.30 and 8.32 into the following equation:
8.34

namely

8.35

Similarly, in the hogging condition, g (t) and H (t) can be obtained as follows:

8.36

8.37

The ultimate moment capacity of the hull under sagging bending moment is

8.38

In the hogging condition, the ultimate moment capacity of the hull is

8.39

To calculate Eq. 8.38 or 8.39, the ultimate strength of the stiffened panel and the
unstiffened plate must be known. Equations 8.5, 8.7, 8.9, 8.20 and 8.21 are used to
predict the ultimate strength of the stiffened panel and the unstiffened plate with
crack damage.

The variance of the ultimate moment capacity of the hull girder under the sagging
condition can be calculated using standard theory. The detailed formulas can be
found in Hu et al. (2004).

> Read full chapter

Lateral-Torsional Buckling
Chai H. Yoo, Sung C. Lee, in Stability of Structures, 2011

7.1 Introduction
A transversely (or combined transversely and axially) loaded member that is bent
with respect to its major axis may buckle laterally if its compression flange is not
sufficiently supported laterally. The reason buckling occurs in a beam at all is that
the compression flange or the extreme edge of the compression side of a narrow
rectangular beam, which behaves like a column resting on an elastic foundation,
becomes unstable. If the flexural rigidity of the beam with respect to the plane
of the bending is many times greater than the rigidity of the lateral bending, the
beam may buckle and collapse long before the bending stresses reach the yield
point. As long as the applied loads remain below the limit value, the beam remains
stable; that is, the beam that is slightly twisted and/or bent laterally returns to its
original configuration upon the removal of the disturbing force. With increasing
load intensity, the restoring forces become smaller and smaller, until a loading is
reached at which, in addition to the plane bending equilibrium configuration, an
adjacent, deflected, and twisted, equilibrium position becomes equally possible. The
original bending configuration is no longer stable, and the lowest load at which
such an alternative equilibrium configuration becomes possible is the critical load
of the beam. At the critical load, the compression flange tends to bend laterally,
exceeding the restoring force provided by the remaining portion of the cross section
to cause the section to twist. Lateral buckling is a misnomer, for no lateral deflection
is possible without concurrent twisting of the section.

Bleich (1952) gives credit to Prandtl (1899) and Michell (1899) for producing the first
theoretical studies on the lateral buckling of beams with long narrow rectangular
sections. Similar credit is also extended to Timoshenko (1910) for deriving the funda-
mental differential equation of torsion of symmetrical I-beams and investigating the
lateral buckling of transversely loaded deep I-beams with the derived equation. Since
then, many investigators, including Vlasov (1940), Winter (1943), Hill (1954), Clark
and Hill (1960), and Galambos (1963), have contributed on both elastic and inelastic
lateral-torsional buckling of various shapes. Some of the early developments of the
resisting capacities of steel structural members leading to the Load and Resistance
Factor Design (LRFD) are summarized by Vincent (1969).

> Read full chapter

Collapse Analysis of Ship Hulls


Yong Bai, Wei-Liang Jin, in Marine Structural Design (Second Edition), 2016

20.3.1 Ultimate Moment Capacity Based on Elastic Section


Modulus
In the initial yield moment approach, it is assumed that the ultimate strength of
the hull girder is reached when the deck (alone) has yielded. Premature buckling is
assumed not to occur. In this approach, the elastic section modulus is the primary
factor for measuring the longitudinal bending strength of the hull. With these
assumptions, the initial yield moment can be written as

(20.27)
here, (SM)e is the elastic section modulus. Owing to the use of greater slenderness
ratios for stiffeners and plate panels, and high yield steels, the possibility of buckling
failure has increased. The initial yield moment may not always be the lower bound
to hull girder strength, since the buckling of the individual structural elements was
not accounted for.

Owing to the simplicity of the initial yield moment equation, it can frequently be
used in practical engineering. Vasta (1958) suggested that the ship hull would reach
its ultimate strength when the compression flange in the upper deck (in the sagging
condition) or the bottom plating (in the hogging condition) collapses, and that the
yield stress in the initial yield moment, Eqn (20.24), may be replaced by the ultimate
strength u of the upper deck or the bottom plating.

Mansour and Faulkner (1973) suggested the Vasta formula can be modified to
account for the shift in the neutral axis location after the buckling of the compression
flange.

(20.28)

where k is a function of the ratio of the areas for a one side shell to the compression
flange. For a frigate, the calculated value of k is approximately 0.1.

Viner (1986) suggested that hull girders collapse immediately after the longitudinal
on the compression flange reaches its ultimate strength, and suggested the follow-
ing ultimate moment equation,

(20.29)

where a is normal in the range of 0.92–1.05 (mean 0.985).

The findings of Mansour and Faulkner (1973) and Viner (1986) are very useful
because of their simplicity—ultimate moment capacity is approximately the product
of the elastic section modulus and the ultimate strength of the compression flange.

Valsgård and Steen (1991) pointed out that hull sections have strength reserves
beyond the onset of the collapse of the hull section strength margin, and suggested
that a is 1.127 for the single-hull VLCC Energy Concentration, which collapsed in
1980.

Faulkner and Sadden (1979) made further modifications,

(20.30)

where U is the ultimate strength of the most critical stiffened panels.

> Read full chapter


The fatigue performance of composite
structural components
M.D. Gilchrist, in Fatigue in Composites, 2003

22.6.6 Bridges and beams


McCormick43 performed an accelerated fatigue test on a glass/polyester pedestrian
footbridge of length 4.9 m. The bridge had been fabricated by bonding three
identical open-web GRP trussed girders to pultruded 12 mm thick GRP plates. A
more severe load-deflection cycle was used to test the bridge than would have been
seen in normal operating conditions. This was done in order to accelerate failure of a
connection or an element of the bridge rather than to evaluate its service life. Visible
cracks in the epoxy adhesive joints between the flange plates and the horizontal
stiffeners were observed at the ends of some of the stiffeners when fatigued for
0.5 × 106 cycles from 7.2 to 17.8 kN at a frequency of 0.15 Hz. This compared with
the post-fatigue monotonic failure load of some 74 kN, which indicates that the
bridge had been over-designed with a factor of safety in excess of four.

Figure 22.12 gives details the fatigue test configuration for pultruded box-beams
and wide flange beams which have been manufactured from unidirectional
E-glass/polyester.44 The wide flange beams exhibited a high resistance to fatigue,
with varying strain ranges up to levels as high as 84% of the ultimate monotonic
strain of the specimen without failing after 6.86 × 106 cycles. Tests on the box beams,
however, with a steel roller as the load-transfer mechanism (i.e. a line load), exhibited
three distinct responses: a fibre-breakage failure mechanism, an endurance limit
and a transition zone. The fatigue limit response was seen during a test with a
strain range of 0.14% (i.e. 18% of the ultimate strain) which did not cause any
damage after 6.9 × 106 cycles. The fibre breakage failure mechanism was seen in
tests at higher strain ranges of 0.18%-0.19% with failure after 4500 cycles. The
transitional response (failure after 0.85 × 106 cycles at an intermediate strain range
of 0.165%) exhibited failure due to a longitudinal matrix crack along the flange-web
intersection.
Fig. 22.12. Three-point flexure testing of box section and wide-flange section
beams.44 (a) Crosssection of test specimens and location of diameter gauges and
strain gauges across width and span. (b) Experimental set-up 1. (c) Experimental
set-up 2.

Meier and co-workers45−47 tested two box-beams in four-point flexure under sinu-
soidal load-control at a frequency of 2 Hz for 108 cycles. The beam flanges were
fabricated from unidirectional profiles surrounded by ± 45° plies, whilst the webs
were filament wound with ± 45° plies of E-glass/epoxy. The interior of the section
was filled with an epoxy-resin foam. Figure 22.13 details the fabrication of these
beams. The maximum loads were 19.1% and 28.6% of the monotonic failure load,
i.e. 40 kN and 60 kN, respectively. The corresponding surface tensile strains at the
outside of the beam were 0.28% and 0.42%. Tests with greater maximum load levels
were not performed. The beam which was subjected to the lowest maximum load did
not develop any detectable damage prior to termination of the test, after 108 cycles.
The bending stiffness of this beam after testing was identical to that before testing.
However, in the more highly loaded beam, matrix cracks initiated on the matrix-rich
surface of the tensile face after 4 × 106 cycles. A zone of delamination (approximately
20 mm × 20 mm in area) developed at 38 × 106 cycles between the outer ± 45° layer
and the unidirectional compression flange at the top corner of the beam under the
loading point. This delaminated area had extended to 20 mm × 28 mm after 50 × 106
cycles. Further cycling to 108 cycles only caused a 2% decrease in the beam bending
stiffness.
Fig. 22.13. (a) and (b) Cross-section of box beams which have been tested under
four-point flexure.47 Dimensions are in mm and the beam length is 2.8 m. A
four-point flexure testing arrangement is used with loads applied at the quarter
points.

The present author and co-workers48−50 have examined the monotonic and cyclic
fatigue failure of composite I beams. Both carbon-fibre/epoxy-matrix and glass-f-
ibre/epoxy-matrix composites were used to manufacture the I beams, which had a
multi-directional stacking sequence consisting of a balanced lay-up of 0, + 45 and
−45° plies. Both un-notched and notched I beams have been studied. A four-point
flexural configuration was used to test the I beams and in no cases were the beams
fatigued above the loads at which buckling of the compression flanges occurred
during monotonic testing. The carbon/epoxy and glass/epoxy I beams which were
un-notched did not exhibit any detectable damage within 1.2 × 106 and 8 × 106
fatigue cycles, respectively. At these numbers of cycles the fatigue tests were halt-
ed. The excellent fatigue behaviour of these un-notched composite I beams was
most noteworthy, especially since the maximum fatigue loads which were applied
represented typically about 75 to 100% of the loads needed to cause buckling of
the compression flange during monotonic testing. Subsequently, the I beams were
notched in order to induce failure under the fatigue loads. The most damaging type
of notch that was introduced was a 60 mm diameter hole in the web section of the I
beam (see Fig. 22.14). The notched I beams then failed under the fatigue loads. For
example, in the case of the carbon-fibre/epoxy I beams which were fatigue loaded
at 5 Hz from 5 kN to 50 kN (which represented about 9 to 90% of the load at
which buckling of the compression flanges occurred during monotonic testing) they
failed after 4.78 × 106 cycles. Fatigue failure was due to various types of damage,
including delamination, matrix microcracking and fibre fracture, occurring around
the 60 mm diameter web notch. The damage mechanisms in the notched CFRP I
beams were studied in detail. The most severe of such damage was caused by the
tensile stresses which were present around the web notch. The principal mode of
damage was matrix cracking, in plies oriented at 90° to the local direct tensile stress.
A significant proportion of this damage occurred within the first 0.5 × 106 cycles.
The matrix cracking led eventually to delamination and fibre fracture, the latter being
the final cause of structural failure of the I beams.
Fig. 22.14(a). I beams manufactured using two channel sections, two flange caps and
strips of wound prepreg tow.48−50 The 24-ply lay-up has a (−45/0/45°)2s(45/0/−45°)2s
global stacking sequence, where the global 0° is along the axis of the I beam
as shown. The corresponding local coordinate system (not shown here) used for
microscopy around web notches is (0/−45/90°)2s (90/−45/0°)2s (cf. Fig. 22.14(b)).

Fig. 22.14(b). Schematic diagram of the four-point flexure loading arrangement


used for flexural fatigue of I beams.48−50 The shear stress around the web notches can
be resolved into direct tensile and compressive stresses which act in four different
quadrants, Tt, Tc, Cc and Ct, as shown. Quadrants Tt and Tc are subject to a resolved
tensile stress state and Cc and Ct to a resolved compressive stress state. Quadrants
Tt and Ct are adjacent to the tension flange whilst Tc and Cc are adjacent to the
compression flange. Sections for optical microscopy were taken after testing from
around the web notches, typically in quadrants Tt and Tc. Such sections were parallel
to the global ± 45° orientation (cf. Fig. 22.14(a)).

> Read full chapter

Example of a Steel Frame Building


Retrofitted with Concentric Braces☆
Abdoreza S. Moghadam, ... Mahdi Eghbali, in Advanced Design Examples of Seismic
Retrofit of Structures, 2019

5.6.1.1 Beams
The strength of structural steel elements under flexural actions shall be calculated
in accordance with this section if the calculated axial load does not exceed 10% of
the axial strength.

For bare beams bent about their major axes and symmetric about both axes, sat-
isfying the requirements of compact sections, Lb < Lp, QCE shall be computed in
accordance with Eq. (5.20):

(5.20)
where:

Lb = distance between points braced against lateral displacement of the compression


flange, or between points braced to prevent twist of the cross section, per AISC 360
[8];

Lp = limiting lateral unbraced length for full plastic bending capacity for uniform
bending from AISC 360;

MpCE = expected plastic moment capacity; and

Fye = expected yield strength of the material.

Based on the Iranian National Building Code; Part 10: Steel Building [6], the flexural
members are considered braced if the free distance of the compression flange is
smaller than the minimum of the results of Eqs. (5.21) and (5.22).

(5.21)

(5.22)

where:

bf = width of the compression flange;

Af = area of the compression flange; and

d = height of the flexural member;

As an example and for IPE 140 we have:

All the beams that carry gravity loads are braced, because the distance of the joists
in the diaphragms is smaller than Lc. For the beams that do not carry gravity loads,
and where the conditions of braced section are not satisfied, the expected flexural
capacity of the section, QCE, shall be computed in accordance with Eq. (5.23):

(5.23)

where S = the elastic section modulus of a member.

The expected flexural capacity of the beams is presented in Table 5.8.

Table 5.8. Expected Flexural Capacity of Beams

MCE (ton m)
Profile Z (cm3) S (cm3) Fye (kg/cm2) Braced Unbraced
2INP240 829 708 2360 19.56 18.38
INP240 411 354 2360 9.70 9.19
2IPE140 172 154.6 2360 4.04 4.01
IPE140 88 77.3 2360 2.07 2.01
IPE180 166 146 2360 3.91 3.79
IPE240 366 324 2360 8.63 8.41

If the beam strength is governed by the shear strength of the unstiffened web and ,
then VCE shall be calculated in accordance with Eq. (5.24):

(5.24)

where:

Aw = nominal area of the web;

tw = web thickness; and

h = distance from inside of compression flange to inside of tension flange.

The expected shear capacity of beams is presented in Table 5.9.

Table 5.9. Expected Shear Capacity of Beams

Profile Fye (kg/cm2) Aw (cm2) VCE (ton)


2INP240 2360 33.4 47.3
INP240 2360 16.7 23.64
2IPE140 2360 10.52 14.9
IPE140 2360 5.26 7.44
IPE180 2360 7.73 10.94
IPE240 2360 11.78 16.68

> Read full chapter

Recent Developments
Xiao-Ling Zhao, ... Gregory Hancock, in Cold-Formed Tubular Members and Con-
nections, 2005

9.3.2 Bolted Moment End Plate Behaviour


Tests were conducted by Wheeler et al (1995) on bolted moment end plate connec-
tions joining RHS. Figure 9.4 shows such a connection under testing. An analytical
model to predict the serviceability limit moment and ultimate moment capacities
of such connections has been presented in Wheeler et al (1998). The connection
geometry considered utilizes two rows of bolts, one of which is located above
the tension flange and the other of which is positioned symmetrically below the
compression flange, as shown in Figure 9.5. Using a so-called modified stub-tee
approach, the model considers the combined effects of prying action caused by
flexible end plates and the formation of yield lines in the end plates as shown
in Figure 9.6. The model has been calibrated against experimental data given in
Wheeler et al (1995). Finite element analysis of such connections was also reported
in Wheeler et al (2000).

Figure 9.4. Bolted moment end plate connections under testing (Wheeler et al 1995)

Figure 9.5. End plate layout and model parameters (Wheeler et al 1998)
Figure 9.6. Yield line mechanisms for bolted moment end plate connection (Wheeler
et al 1998)

Of the three types of end plate behaviour considered in the stub-tee model (thick,
thin and intermediate), Wheeler et al (1998) recommended that the end plate
connections be designed to behave in an intermediate fashion, with the connec-
tion strength being governed by tensile bolt failure. Thin plate behaviour results
in connections that are more ductile and exhibit extremely high rotations, while
connections exhibiting thick plate behaviour have much less rotation capacity and
may be uneconomical.

> Read full chapter

Direct strength method—a general ap-


proach for the design of cold-formed
steel structures
D. Camotim, ... A.D. Martins, in Recent Trends in Cold-Formed Steel Construction,
2016

4.1.2 Historical perspective


The designation “Direct Strength Method” was first mentioned in the pioneering
work of Schafer and Peköz (1998a,b), in the context of developing new design
approaches for CFS beams. In particular, the authors explored the application of
strength curves to provide a direct way to assess (1) distortional resistance of
lipped channel and Z-section beams, and (2) the local and distortional resistance of
deck sections with multiple longitudinal intermediate stiffeners in the compression
flange. The method had its roots in the design method proposed earlier by Hancock
et al. (1994) when investigating the distortional resistance of thin-walled sections
under compression or bending.1 These authors (1) adopted the effective section
approach, (2) adapted Winter's expression/curve to distortional buckling, and (3)
modified its coefficients to obtain more accurate estimates. The end product of
this research effort was

[4.1]

where be is the effective part of the wall width b. The width reduction given by
Eq. [4.1] was performed for all cross-section walls, and the method was included
in the 1996 edition of the Australian/New Zealand Standard for Cold-Formed Steel
Structures (AS/NZS, 1996).

A couple of years later, Schafer and Peköz (1998a,b), after going through Hancock
et al. (1994), proposed a modification of Eq. [4.1] which consisted of replacing its
coefficients (0.6/0.25) by new/reduced ones (0.4/0.15). They showed that their
proposal provided, in general, accurate and reliable predictions of the ultimate
strength of CFS beams failing in distortional modes, clearly outperforming the
design procedure prescribed by the AISI specification at the time. Mainly due to the
research carried out by Schafer at Johns Hopkins University (Schafer, 2002a,b, 2006),
this first application of the DSM was fairly quickly followed by similar applications for
(1) CFS columns failing in local, distortional, and global modes, and (2) CFS beams
failing in local and global modes. Such research efforts led to the so-called “first
generation of DSM design curves,” soon to be codified in North America (AISI, 2004)
and Australia and New Zealand (AS/NZS, 2005), which are addressed in detail in
Section 4.2.

The potential of the DSM approach for the design of thin-walled (not necessarily
CFS) members and structural systems was quickly recognized worldwide, and led to
a plethora of experimental and numerical investigations aimed at the development,
validation, and possible codification of DSM-based design methodologies for a wide
variety of structural problems. It may easily be argued that the “frontrunners” of this
“race,” in the sense that the corresponding design curves had already been codified
in North America (AISI, 2012), were the studies concerning CFS (1) beams failing
due to shear or combined bending and shear, advocated by Pham and Hancock
at the University of Sydney, and (2) perforated columns and beams failing in local,
distortional, and global modes, investigated by Moen and Schafer at Johns Hopkins
University—these two design approaches are addressed in some detail in Section
4.3. But the “race” is full of “competitors,” as is clearly evidenced in Section 4.4,
which includes a selection of recent or ongoing studies carried out with the final
goal of achieving DSM-based design expressions/curves for various problems. It is
worth noting that (1) this selection is by no means exhaustive (the main purpose
of Section 4.4 is to illustrate the power and versatility of the DSM approach) and
necessarily reflects the authors' research activity and interests; also, due to space
limitations and the focus of this book, (2) only applications to CFS members and
structural systems are considered. Nevertheless, the reader should be aware that
DSM solutions to problems involving other materials are currently being sought.

It should be pointed out that the DSM bears some resemblance (1) to the “general
method” prescribed in Part 1–1 of Eurocode 3 (EC3-1–1 (CEN, 2005)) for the design
of structural components (members or plane frames/subframes) against lateral
and lateral–torsional failures, and (2) to the application of the “Overall Interaction
Concept,” an ambitious endeavor championed by Boissonnade et al. (2013) that
circumvents the need to perform a cross-section classification procedure (currently
unavoidable when designing according to Eurocode 3). Finally, one must mention
the “Continuous Strength Method” (CSM), devised and developed mainly due
to the efforts of Gardner and associates at Imperial College London (Gardner,
2008; Gardner et al., 2013; Afshan and Gardner, 2013), and also not requiring any
cross-section classification. Although the CSM and DSM philosophies are different,
they share some features and may be viewed as alternative design approaches to
some structural problems. The CSM was initially proposed in the context of the
design of stainless steel (nonlinear stress–strain relationship) members subject to
compression or bending against local failures (Ashraf et al., 2006) and later extended
to cover high-strength and carbon steel members (Gardner, 2008). The applica-
tion of the method hinges on an experimentally derived “base curve” relating the
cross-section resistance and deformation capacity, with the latter determining the
cross-section ability to evolve into the strain-hardening region, thus sustaining a
higher loading—naturally, this design approach makes it possible to take advantage
of the added strength due to strain-hardening. An additional benefit of the CSM is
the fact that it readily and explicitly provides ductility information.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like