You are on page 1of 30

34

Fatigue and Fracture

34.1 Introduction .................................................. 34-1


34.2 Fatigue ......................................................... 34-2
Fatigue Resistance  Fatigue Loading  Categorization of
Details  Low-Cycle Fatigue and Seismic Frame Details
34.3 Fracture ........................................................ 34-15
Fracture Resistance  Material Specification and Detailing
Robert J. Dexter to Avoid Fracture
Department of Civil Engineering,
University of Minnesota, 34.4 Summary ...................................................... 34-24
Minneapolis, MN References ............................................................. 34-24

34.1 Introduction
Fatigue and fracture are two important limit states that need to be checked in the design and evaluation
of structures. Fatigue is the formation of a crack due to cyclic loading. The fatigue limit state is defined as
the development of a through-thickness crack. This is a serviceability limit state and does not necessarily
mean that the structure is in danger of fracture or collapse. Fracture is the rupture in tension or rapid
extension of a crack leading to gross deformation, loss of function or serviceability, or complete
separation of the component.
This section of the handbook presents an overview of information useful to structural engineers in
evaluating the fatigue and fracture limit states of steel, aluminum, and concrete structural components.
Topics include materials selection, design, and detailing for new structures, as well as assessment of
existing structures. The emphasis of this chapter is on structural steel components, since aluminum and
other metal components are not common in the primary load-carrying systems of most civil structures.
Fatigue of concrete components is covered only briefly since it is rarely a significant problem. As a
practical matter, fracture of concrete is checked by usual strength design calculations and therefore is
not covered here. The fracture mechanics of concrete is covered elsewhere [1].
Although most of the examples involve buildings or bridges, the information is equally applicable
to similar details in cranes, ships, offshore structures, heavy vehicle frames, etc. The primary difference
between various structure types is in the applied loading, whereas the type of structure does not affect
the resistance of the details to fatigue or fracture.
Since the scope of this section is limited to practical information, there are many interesting aspects of
fatigue and fracture that are not discussed. There are several good texts that can serve as a starting point
for more in-depth studies [2–4].

0-8493-1569-7/05/$0.00+$1.50
# 2005 by CRC Press 34-1

Copyright 2005 by CRC Press


34-2 Handbook of Structural Engineering

34.2 Fatigue
Fatigue resistance for a particular detail is the allowable constant-amplitude stress range (S) for a
specified number of cycles of loading (N), typically expressed in an S–N curve. S–N curves are based on
fatigue tests of full-scale members with welded or bolted details. The assignment of various details
to specific categories for S-N curve analysis is discussed below, including bolts, anchor rods, hollow
structural sections, concrete members, and cables.
When information about a specific crack is available, a fracture mechanics crack growth rate analysis
should be used to calculate the remaining life [3,4]. However, in the design stage, without specific initial
crack size data, the fracture mechanics approach is not any more accurate than the S–N curve approach
[5]. Therefore, the fracture mechanics crack growth analysis will not be discussed further.

34.2.1 Fatigue Resistance


The approach to designing and assessing structures for fatigue is empirical and is based on tests of
full-scale members with welded or bolted details. Such tests indicate that
 The strength and type of steel have only a negligible effect on the fatigue resistance expected for
a particular detail [6–10].
 The welding process also does not typically have an effect on the fatigue resistance [11,12].
 The primary effect of constant-amplitude fatigue loading can be accounted for in the live-load
stress range [6–9], that is, the mean stress is not significant.
The independence of fatigue resistance from the type of steel greatly simplifies the development of
design rules for fatigue since it eliminates the need to generate fatigue data for every type of structural
steel. Testing that has been performed on welded details in stainless steel show that the fatigue strength of
stainless steel details is comparable to that of ferritic structural steel details [13].
The reason that the dead load has little effect is that, locally, there are very high residual stresses.
In details that are not welded, such as anchor rods, there is a mean stress effect [14]. A worst-case
conservative assumption, that is, a high tensile mean stress, is made in the testing and in the design of
these nonwelded details.

34.2.1.1 Standard S–N Curves


The nominal stress approach is the established approach for fatigue design and assessment of steel and
aluminum structures. The nominal stress approach is based on S–N curves, where S is the nominal stress
range and N is the number of cycles until the appearance of a visible crack. Details are designed based on
the nominal stress range in the connecting members rather than the local ‘‘concentrated’’ stress at the
detail. The nominal stress is usually obtained from standard design equations for bending and axial stress
and does not include the effect of stress concentrations of welds and attachments. Usually, the nominal
stress in the members can be easily calculated without excessive error. However, the proper definition of
the nominal stresses may become a problem in regions of high stress gradients.
Figure 34.1 shows the S–N curves for steel that are used throughout North America for a variety of
welded structures, including the American Association of State Highway and Transportation Officials
(AASHTO) bridge design specifications [15], the American Institute for Steel Construction (AISC)
Manual of Steel Construction [16], the American Railway Engineering and Maintenance-of-Way
Association (AREMA) Manual for Railway Engineering [17], the American Welding Society (AWS D1.1)
Structural Welding Code [18], and the Canadian Standards Association (CSA S16–2001) Limit States
Design of Steel Structures [19].
AASHTO has seven S–N curves for seven categories (A through E0 ) of weld details. The fatigue design
procedure is based on associating the weld detail under consideration with a specific category. The effects
of the welds and other stress concentrations, including the typical defects and residual stresses, are
reflected in the ordinate of the S–N curves for the various detail categories.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-3

AASHTO curves
1000
100

Stress range, MPa

Stress range, ksi


A
B
100 B′
C 10
D
E

E′

10
104 105 106 107 108
Number of cycles

FIGURE 34.1 S–N curves for the seven primary fatigue categories from the AASHTO, AREMA, AWS, and AISC
specifications; the dotted lines are the constant-amplitude fatigue limits (CAFLs) and indicate the detail category.

TABLE 34.1 Parameters for S –N Curves


AASHTO,
AISC, and Coefficient A CAFLa for Equivalent Nearest Coefficient A CAFL for
AWS for steel steel Eurocode 3 BS 7608 for aluminum Exponent aluminum
category (MPa3) (MPa) category category (MPa3) for aluminum (MPa)

A .81.9  1011 165 160 B 2.18  1019 6.85 70


B .39.3  1011 110 125 C 1.95  1014 4.84 37
B0 .20.0  1011 83 100 D N/A N/A N/A
C .14.4  1011 69 90 D 8.88  1011 3.64 28
D .7.21  1011 48 71 F 2.08  1011 3.73 17
E .3.61  1011 31 56 G 3.14  1011 3.45 13
E0 .1.28  1011 18 40 W N/A N/A N/A
a
CAFL, constant-amplitude fatigue limit.

The inverse slope of the regression line fit to the test data for welded details is typically in the range
2.9 to 3.1 [8]. In the AISC, AASHTO, AREMA, and CSA codes as well as in Eurocode 3 (EU3) [20], the
slopes have been standardized at 3.0. Therefore, the S–N curves can be represented by the following
power–law relationship:
A
N¼ ð34:1Þ
S3
where N is the number of stress cycles, S is the nominal stress range, and A is a constant particular to the
detail category, as given in Table 34.1.
AISC and AWS also have a Category F with a slope greater than 3 for checking the shear stress in the
throat of welds. AASHTO and CSA S16 require checking the shear stress in the throat of welds according
to Category E rather than Category F. However, the minimum weld size requirements and the weld
strength requirements generally will ensure sufficient weld throat to avoid fatigue. There are few, if any,
documented fatigue cracking cases associated with shear stress through the throat of fillet welds,
therefore it is unnecessary to check the shear stress range in the throat of welds for fatigue and this will
not be discussed further.
Eurocode 3 (EC3)0 [20] and the British Standard 7608 [21] also use a nominal stress approach, but
they each have unique sets of S–N curves with different category labels. However, the end result of

Copyright 2005 by CRC Press


34-4 Handbook of Structural Engineering

checking the stress range will be approximately the same regardless of which of these sets of S–N curves is
used, since all of these design specifications are empirical and are calibrated with the same database of
full-scale test results. None of these sets of S–N curves has any inherent advantages or will estimate the
fatigue life of a detail more accurately than any other set of S–N curves. Seven of the fourteen Eurocode
curves are the same as the seven AASHTO curves, as indicated in Table 34.1. The American Bureau of
Shipping (ABS) guidelines [22], the U.K. Health and Safety Executive [23], and other groups in the
marine industry use S–N curves from the British Standards (BS 7608) [21]. The BS 7608 S–N curves can
be associated approximately with the AASHTO S–N curves so that knowledge of the fatigue strength of
details in this specification can be translated to an equivalent AASHTO category. Table 34.1 cross-
references the BS 7608 S–N curves with the nearest AASHTO S–N curves.
In the nominal stress range approach, each detail category has a constant-amplitude fatigue limit
(CAFL), which is also given in Table 34.1. The CAFL is a stress range below which no fatigue cracks
occurred in tests conducted with constant-amplitude loading. These limits are shown as the horizontal
lines on the right side of the S–N curves in Figure 34.1. The Eurocode and BS S–N curves define the
fatigue limits differently.

34.2.1.2 The Effect of Corrosion of Fatigue Resistance


The full-scale fatigue experiments on which the fatigue rules are based were carried out in moist air and
therefore reflect some degree of environmental effect or corrosion fatigue. Therefore, the lower-bound
S–N curves in Figure 34.1 can be used for the design of details with a mildly corrosive environment (such
as the environment for bridges, even if salt or other corrosive chemicals are used for deicing) or provided
with suitable corrosion protection (galvanizing, other coating, or cathodic protection). Some design
codes for offshore structures have reduced fatigue life (by approximately a factor of 2) when details are
exposed to seawater [21,23].
At relatively high stress ranges at which most accelerated tests are conducted, the effect of seawater is
clearly detrimental. However, there is evidence that the effect of corrosion in seawater is not so severe for
long-life variable-amplitude fatigue of welded details. Full-scale fatigue experiments in seawater at
realistic service stress ranges do not show significantly lower fatigue lives [24], provided that corrosion is
not so severe that it causes pitting or significant section loss. At relatively low stress ranges near the
CAFL, typical of service loading, it appears that build-up of corrosion product in the crack may increase
crack closure and retard crack growth, at least enough to offset the increase that would otherwise occur
due to the environmental effect. The fatigue lives seem to be more significantly affected by the stress
concentration at the toe of welds than by the corrosive environment.
Severely corroded members may be evaluated as Category E details [25], regardless of their original
category (unless of course they were Category E0 or worse to start with). However, pitting or significant
section loss from severe corrosion can lower that fatigue strength [25,26].

34.2.1.3 S–N Curves for Aluminum


S–N curves have been proposed by the Aluminum Association for welded aluminum structures [27,28].
The categories for details are similar to those for steel, although the fatigue strength of a detail is
approximately 38% of the fatigue strength of the same detail in steel. For aluminum, the exponent in
Equation 34.1 is variable and is given in Table 34.1 with the coefficient A and CAFL. The design
procedures are similar and the classification of details into categories is approximately the same.

34.2.2 Fatigue Loading


Since fatigue is typically only a serviceability problem, members and connections are designed for fatigue
using service loads. Most structures experience what is known as long-life variable-amplitude loading,
that is, very large numbers of random-amplitude cycles greater than the number of cycles associated with
the CAFL [29]. For example, a structure loaded continuously at an average rate of once per minute
(0.016 Hz), would accumulate 10 million cycles in 19 years. With long-life variable-amplitude loading,

Copyright 2005 by CRC Press


Fatigue and Fracture 34-5

structures are usually designed so that only a small fraction of cycles, on the order of 0.01%, exceed the
CAFL [29,30].
If the percentage of stress ranges exceeding the CAFL is greater than 0.01%, the history of N variable
stress ranges can be converted to N cycles of an effective stress range that can then be used just like the
constant-amplitude stress range in S–N curve analysis. Typically, Miner’s Rule [31] is used to calculate
an effective stress range from a histogram of variable stress ranges. Theoretically, this effective constant-
amplitude stress range results in approximately the same fatigue damage for a given number of cycles as
that for the same number of cycles of the variable-amplitude service history. If the stress ranges are
counted in discrete ‘‘bins,’’ as in a histogram, the effective stress range, SRe [29] can be calculated as
X 1=3
SRe ¼ ðai Sri3 Þ ð34:2Þ
i

where ai is the number of stress cycles with stress range in the bin with average value Sri divided by the
total number of stress cycles (N ).
Variable-amplitude fatigue tests conducted with various sequences in the variable-amplitude loading
history have shown that Miner’s Rule is reasonably accurate in most cases but can be unconservative
with some load histories with unusual sequences. For this reason, some fatigue design specifications for
offshore structures put a safety factor of 2.0 on life if Miner’s Rule is used [23].
In the AASHTO specifications [15], the stress range from the fatigue design truck represents the
effective stress range. No additional safety factor is used for Miner’s Rule since it is relatively accurate for
truck loading on bridges. For large numbers of cycles, the AASHTO specification has another check that
involves comparing the stress range from the fatigue design truck to half of the CAFL. The rationale for
this check is that if the effective stress range is less than half the CAFL, most of the stress ranges should be
below the CAFL, but occasionally (about once a day) the stress range can exceed the CAFL with no
significant effect.
Misalignment at a welded joint is a primary factor in susceptibility to cracking. The misalignment
causes eccentric loading, local bending, and stress concentration. The stress concentration factor (SCF)
associated with misalignment is
SCF ¼ 1:0 þ 6e=t ð34:3Þ
where e is the eccentricity and t is the smaller of the thicknesses of two opposing loaded members.
The nominal stress times the SCF should then be compared to the appropriate category. Generally,
such misalignment should be avoided at fatigue critical locations. Equation 34.3 can also be used where
e is the distance that the weld is displaced out of plane due to angular distortion. A thorough guide to
the SCF for various types of misalignment and distortion, including plates of unequal thickness, can be
found in British Standard BS 7910 [32].

34.2.3 Categorization of Details


The following is a brief simplified overview of the categorization of fatigue details. In all cases, the
applicable specifications should also be checked. Several reports have been published that show a large
number of illustrations of details and their categories in addition to those in AISC and AASHTO
specifications [33,34]. Also, the EC3 [20] and the British Standard BS 7608 [21] have more detailed
illustrations for their categorization than AISC or AASHTO specifications. A book by Maddox [9]
discusses categorization of many details in accordance with BS 7608, from which roughly equivalent
AASHTO categories can be inferred (using Table 34.1). In most cases, the fatigue strength recommended
in these European standards is similar to the fatigue strength in the AISC and AASHTO specifications.
However, there are several cases where the fatigue strength is significantly different; usually, the
European specifications are more conservative.
There have been very few, if any, failures that have been attributed to details with fatigue strength
greater than Category C. Most structures have many Category C or even more severe details, and these

Copyright 2005 by CRC Press


34-6 Handbook of Structural Engineering

will generally govern the fatigue design. Therefore, only Category C and more severe details will be
discussed in depth.

34.2.3.1 Welded Joints


Welded joints are considered transverse if the axis of the weld is perpendicular to the primary stress
range. Unless there is a stress concentration from the configuration of the detail, transverse welds are
typically Category C details. For example, the transverse welds connecting stiffeners to the web and
flange of girders are Category C details.
Full-penetration groove welded butt joints subjected to nondestructive evaluation (NDE) such as
ultrasonic testing or radiographic testing are also Category C details. If the reinforcement of these
full-penetration butt joints is ground smooth, they are Category B details. Tests show that groove welds
that contain large internal discontinuities that were not screened out by NDE had fatigue strength
comparable to Category E [5]. BS 7608 [21] and BS 7910 [32] have reduced fatigue strength curves for
groove welds with defects that are generally in agreement with these experimental data. Transverse
groove welds with a permanent backing bar are reduced to Category D [35,36]. One-sided welds with
melt through (without backing bars) are also classified as Category D [35,36].
Continuous longitudinal welds are Category B or B0 details. However, the termination of a long-
itudinal fillet weld is more severe (Category E). The termination of full-penetration groove longitudinal
welds requires a ground transition radius but gives greater fatigue strength, depending on the radius.
If longitudinal welds must be terminated, it is better to extend the welds to a location along the structural
member where the stress ranges are small or entirely in compression.
Cope holes, cutouts, and snipes should be used for weld access and to avoid intersecting welds. If the
thermally cut surfaces have edges conforming to the ANSI smoothness of 1000, these may be considered
Category D details, even if welds terminate at these details [35,36]. A rougher cope hole may be treated as
a Category E detail. If the steel at the thermally cut edge transforms to martensite, in some cases small
cracks may occur that will propagate at even lower stress ranges.
Attachments normal to flanges or plates that do not carry significant load are rated Category C if less
than 51 mm long in the direction of the primary stress range, D if between 51 and 101 mm long, and E if
greater than 101 mm long. If there is not at least a 10-mm edge distance, then Category E applies for an
attachment of any length. Category E0 , slightly worse than Category E, applies if the attachment plates
or the flanges exceed 25 mm in thickness.
The cruciform joint where the load-carrying member is discontinuous is considered a Category C
detail because it is assumed that the plate transverse to the load-carrying member does not have any
stress range. A special reduction factor for the fatigue strength is provided when the load-carrying
plate exceeds 13 mm in thickness. This factor accounts for the possible crack initiation from the unfused
area at the root of the fillet welds (as opposed to the typical crack initiation at the weld toe for thinner
plates) [37].
In most other types of load-carrying attachments, there is interaction between the stress range in
the transverse load-carrying attachment and the stress range in the main member. In practice, each of
these stress ranges is checked separately. The attachment is evaluated with respect to the stress range
in the main member, and then it is separately evaluated with respect to the transverse stress range.
The combined multiaxial effect of the two stress ranges is taken into account by a relative decrease in
the fatigue strength, that is, most load-carrying attachments are considered Category E details.

34.2.3.2 Bolted Joints and Anchor Rods


Small holes are considered Category D details. Therefore, rivetted and mechanically fastened joints
(other than high-strength bolted joints) loaded in shear are evaluated as Category D in terms of the
net-section nominal stress. High-strength A325 and A490 bolted joints that are properly pretensioned
and loaded in shear are Category B details in terms of the net-section nominal stress. Pin plates and
eyebars are designed as Category E details in terms of the net-section nominal stress.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-7

Bolted joints loaded in direct tension are more complicated. Typically, these provisions are applied to
hanger-type or bolted flange connections where the bolts are pretensioned against the plies. In this case,
the total fluctuating load is resisted by the area of the precompressed plies, so that the bolts are subjected
to only a fraction of the total load [38]. However, the analysis to determine this fraction is difficult.
In the AISC Specifications [16] and in BS7608 [21], the designer may assume the load range in the bolts
is 20% of the total applied service load (dead plus live load). The total applied service load must include
any prying load. Prying is very detrimental to fatigue, so it is best to minimize prying forces by using
sufficiently thick plates [16].
In the AISC Specifications [16], the stress range in the bolts is calculated on the tensile stress area,
At, given by
 
p 0:9743 2
At ¼ db  ð34:4Þ
4 n
where db is the nominal diameter (the body or shank diameter) and n is threads per inch. (Note that the
constant would be different if metric threads are used.)
Test data on bolts in direct tension (with the stress range computed on the tensile stress area) are
shown in Figure 34.2 [39,40]. The M20 bolts and B7 rods are similar in strength to A325 high-strength
bolts (862-MPa tensile strength). Different size diameters of the B7 bolts were also tested, with the results
showing that size did not have an impact on the fatigue limit. Category E seems to be the lower bound
for bolts in direct tension.
Anchor rods in concrete cannot be adequately pretensioned and therefore do not behave like hanger-
type or bolted flange connections. In the double-nut configuration, they are pretensioned between nuts
on either side of the base plate, but the part below the bottom nut is still exposed to the full load range.
Fatigue test data are available for anchor rods [14,41,42]. The data for anchor rods with a double-nut
configuration tightened one third of a turn past snug tight are shown in Figure 34.3. These data agree
well with the bolt data shown in Figure 34.2, that is, the lower-bound fatigue strength is Category E.
These test data were obtained at low enough stress ranges to allow for definition of the CAFL. There
was one failure at 69 MPa but below this stress range there were all runouts. Therefore, the CAFL is much
greater than the CAFL for other Category E details, which is 31 MPa. Choosing from among the CAFLs
for the other categories, the CAFL was taken as that for Category D or 48 MPa.

1000
B7 bolts, Kuperus [39]
100
M20q10.9, Kuperus [39]
Indicates no crack
Stress range, MPa

Stress range, ksi

100
10
D

E′

10
104 105 106 107 108
Number of cycles

FIGURE 34.2 S–N curves for B7 and M20 snug-tight bolts in tension.

Copyright 2005 by CRC Press


34-8 Handbook of Structural Engineering

1000
Frank [41]
Dusel et al. [42] 100
Current
Stress range, MPa Indicates no crack

Stress range, ksi


100
C
10
D
E

E′

10
104 105 106 107 108
Number of cycles

FIGURE 34.3 S–N curves for double-nut anchor rods tightened one third of a turn past snug tight.

1000
Frank [41]
Dusel et al. [42] 100
Current (concentric)
Current (misaligned)
Indicates no crack
Stress range, MPa

Stress range, ksi


100
C 10
D
E

E′

10
104 105 106 107 108
Number of cycles

FIGURE 34.4 S–N curves for snug-tight anchor rods.

Figure 34.4 shows fatigue test data for anchor rods that were only snug tight. Specimens with a 1:40
misalignment are also included. Under these conditions, the lower-bound fatigue strength corresponds
to Category E0 . Since no failures occurred below a stress range of 62 MPa in this case, the CAFL at 48
MPa applies regardless of tightening or alignment.
Based on these data, in the AISC Specifications [16] the fatigue resistance of bolts or anchor rods in direct
tension, on the tensile stress area, is taken as Category E0 in the finite-life range with a CAFL of 48 MPa.

34.2.3.3 Hollow Structural Sections


There have been numerous publications regarding the fatigue strength of rectangular and circular hollow
structural sections (HSS), as summarized in Van Wingerde et al. [43]. The majority of recent work
has taken place in Europe under the advisement of the European Coal and Steel Community (ECCS),
the British Department of Energy (DEn), now called the Health and Safety Executive, and the Comité

Copyright 2005 by CRC Press


Fatigue and Fracture 34-9

International pour le Developement et l’Etude de la Construction Tubulaire (CIDECT) [44]. The


research has resulted in design recommendations that have been adopted by the International Institute
of Welding (IIW Document XIII-1804-99 and XV-1035-99) [45] and CIDECT Design Guide #8 [46] and
have been proposed for inclusion into AWS D1.1 [18] and EC3 [20]. The CIDECT fatigue design
guidelines include both the nominal-stress approach and the hot-spot stress approach. AWS D1.1 [18]
also has nominal-stress and hot-spot stress S–N curves for HSS, although these are not considered as
accurate as the CIDECT guidelines. Unlike AWS, the CIDECT guidelines distinguish between circular
hollow sections (CHS) and rectangular hollow sections (RHS).
Many sign, signal, and light supports are fabricated from HSS. AASHTO has published a new
specification for these types of structures [47] that includes fatigue design using the nominal-stress
approach and classification of the common details with respect to the AWS S–N curves, based on
research in NCHRP Report 412 [14]. In addition, this specification has fatigue design loads for natural
wind gusts, galloping, and vortex shedding. Many aspects of this specification could also be applied to
wind turbines, communication towers, and other flexible structures exposed to wind.
The hot-spot stress approach is similar to the nominal-stress approach and is also based on full-scale
tests, except that the S–N curves in this approach are based on the hot-spot stress range [48]. The hot-
spot stress is the stress normal to the weld axis, originally defined at some small distance (about 6 mm)
from the weld toe. The hot-spot stress can be measured, calculated using finite-element analysis, or
determined from parametric formulas derived from the finite-element analyses and measurements.
A distinction is made between the geometric stress concentration, that is, that associated with the
arrangement of the members, and the local stress concentration, that is, that associated with the weld toe.
In the hot-spot approach, the effect of the geometric stress concentration, expressed as the SCF, is taken
out of the fatigue resistance and instead is included in the analysis. This has the advantage of collapsing
all the S–N curves for different categories into a single baseline S–N curve, but it has the disadvantage of
increasing the complexity of the analysis. A simple structural analysis is usually performed to obtain the
axial and bending forces entering a joint, from which nominal stresses are calculated. The hot-spot stress
at a weld is determined by multiplying the appropriate nominal stress by the SCF. The test data and the
baseline S–N curve still include the effect of the local stress concentration, which is impossible to
calculate accurately and therefore must still be treated empirically.
A simple approach that works well is to define the hot-spot stress as the stress measured with a 3-mm
strain gage placed as close as practically possible to the weld toe, that is, centered about 6 mm from the
weld toe [49]. This is essentially the definition originally used by AWS [48]. The baseline S–N curve used
with the hot-spot stress defined this way is essentially the same as the nominal-stress S–N curve
(Category C) for a transverse butt or fillet weld in a nominal membrane stress field, that is, a stress field
without any geometric stress concentration [49]. This makes sense since the stress at the weld toe of
this detail would include the local stress concentration but would not be affected by geometric stress
concentration. In other words, the SCF is equal to 1.0, and the hot-spot stress is equal to the nominal
stress in this detail.
The CIDECT research updated parametric equations used for SCF calculations and unified the
definition of hot-spot stress. The definition of hot-spot stress is more complex than the simple AWS
definition and involves extrapolating the stress from multiple strain gage measurements or analysis
points [46]. However, the CIDECT design guidelines contain parametric equations that can be used to
calculate SCF based on nondimensional parameters (brace to chord diameter ratio, brace diameter to
thickness ratio, brace to chord thickness ratio, and chord length to diameter ratio) for different types of
loading. A minimum SCF of 2.0 is recommended.
Obviously, the SCF and the baseline curve are dependent on the definition of the hot-spot stress [50].
The CIDECT baseline hot-spot S–N curve is the T0 curve from BS 7608 (equivalent to the Eurocode class
114 curve) for use with 16-mm-thick sections (see Figure 34.5). This is about 25% higher than the baseline
S–N curve used with the AWS-defined hot-spot stress (6 mm from the weld toe). In hollow structural
joints, there is a pronounced thickness effect, with thinner sections having higher fatigue lives. CIDECT
included the thickness effect in the definition of the hot-spot S–N curves, as shown in Figure 34.5.

Copyright 2005 by CRC Press


34-10 Handbook of Structural Engineering

1000
100

Hot-spot stress range, MPa

Hot-spot stress range, ksi


100
t = 4 mm
t = 5 mm
t = 8 mm
10
t = 12 mm
t = 16 mm
t = 25 mm
t = 32 mm
t = 50 mm

10
104 105 106 107 108 109
Number of cycles to failure, Nf

FIGURE 34.5 Hot-spot S–N curves for CHS joints (4 mm t 50 mm) and RHS joints (4 mm t 16 mm).

34.2.3.4 Reinforced and Prestressed Concrete, Strands, and Cables


Because of the relatively low stresses, concrete structures are typically less sensitive to fatigue than welded
steel and aluminum structures. However, fatigue may govern the design when impact loading is
involved, such as pavement, bridge decks, and rail ties. Also, in high-strength concrete the applied stress
ranges increase, and so should the concern for fatigue.
According to ACI Committee Report 215R-74 in the Manual of Standard Practice [51], the fatigue
strength of plain concrete at 10 million cycles is approximately 55% of the ultimate strength. However,
even if failure does not occur, repeated loading may contribute to premature cracking of the concrete,
such as inclined cracking in prestressed beams. This cracking could then lead to localized corrosion and
fatigue of the reinforcement [52].
The fatigue strength of straight, unwelded reinforcing bars and prestressing strand can be described
(in terms of the categories for steel details described earlier) with the Category B S–N curve. The lowest
stress range that has been known to cause a fatigue crack in a straight reinforcing bar was 145 MPa,
which occurred after more than a million cycles. Mean stress level, yield strength, bar size, geometry, and
deformations had minimal effect. ACI Committee 215 suggests that members be designed to limit the
stress range in the reinforcing bar to 138 MPa for high levels of mean stress (increases up to 160 MPa
as mean stress is reduced).
Fatigue tests show that previously bent bars had only about half the fatigue strength of straight bars
and failures have occurred down to 113 MPa [53]. Committee 215 recommends that half of the stress
range for straight bars be used, that is, 69 MPa for the worst-case mean stress. Equating this recom-
mendation to the S–N curves for steel details, bent reinforcement may be treated as a Category D detail.
Provided the quality is good, butt welds in straight reinforcing bars do not significantly lower the
fatigue strength. However, tack welds reduce the fatigue strength of straight bars about 33%, with
failures occurring as low as 138 MPa. Fatigue failures have been reported in welded wire fabric and bar
mats [54].
Batchelor et al. [55] tested deck panels with a single stationary concentrated load applied at the center.
The contact area represented the assumed contact area of the pneumatic tires of large trucks. The only
fatigue tests that did not result in failure at the end of 2.5 and 3 million cycles were performed with a
loading of less than the 50% of the estimated ultimate punching strength of the deck (Pu). Other research
also supports a fatigue limit of 0.5Pu at 2 to 3 million cycles for a stationary pulsating load [56,57].

Copyright 2005 by CRC Press


Fatigue and Fracture 34-11

However, the behavior and the fatigue limit are different in the case of rolling loads, as on a bridge.
Flexural cracking along the reflection of the longitudinal and transverse reinforcement on the bottom
surface was detected in bridges in service, and this is different from cracking patterns in tests with
stationary pulsating loads [58]. Tests with moving or rolling loads indicate that the fatigue limit was as low
as 0.21Pu at 2 to 3 million cycles, comparable to the average flexural cracking load level of about 0.26Pu
[58–60]. In some of the tests [56,58,59] it was demonstrated that the transverse cracks from the con-
structional period and the water penetration during service life decreased the ultimate punching shear and
fatigue strengths of the reinforced concrete deck. However, these studies have not established any quan-
titative interaction between the deterioration from the environmental factors and the repetitive axle load.
Test data show that measured stress ranges in the reinforcement at the location of cracks in a highly
deteriorated deck under high axle loads are less than 35 MPa, well below the 138-MPa threshold
discussed earlier [61]. Therefore, under service load, fatigue does not appear to be a problem for deck
reinforcement. This is consistent with the fact that fatigue of reinforced concrete decks is governed by
punching failure of the concrete part of the structure.
If prestressed members are designed with sufficient precompression so that the section remains
uncracked, there is not likely to be any problem with fatigue. This is because the entire section is resisting
the load ranges, and the stress range in the prestessing strand is minimal. Similarly, for unbonded
prestessed members, the stress ranges will be very small. However, there is reason to be concerned about
bonded prestressing at cracked sections because the stress range increases locally. The concern for
cracked sections is even greater if corrosion is involved. The pitting from corrosive attack can drama-
tically lower the fatigue strength of the reinforcement [52].
Although the fatigue strength of the prestressing strand in air is about equal to Category B, when the
anchorages are tested as well, the fatigue strength of the system is as low as half the fatigue strength of the
wire alone (i.e., about Category E). When actual beams are tested, the situation is very complex, but it is
clear that much lower fatigue strength can be obtained [62,63]. Committee 215 has recommended the
following for prestressed beams:
1. The stress range in prestressed reinforcement, determined from an analysis considering the section
to be cracked, shall not exceed 6% of the tensile strength of the reinforcement. (Author’s note: this
is approximately equivalent to Category C.)
2. Without specific experimental data, the fatigue strength of unbonded reinforcement and its
anchorages shall be taken as half of the fatigue strength of the prestressing steel. (Author’s note:
this is approximately equivalent to Category E.) Lesser values shall be used at anchorages with
multiple elements.
The Post-Tensioning Institute (PTI) has issued Recommendations for Stay Cable Design, Testing, and
Installation [64]. The PTI recommends that uncoupled bar stay cables are Category B details, while
coupled (glued) bar stay cables are Category D. The fatigue strengths of stay cables are verified through
fatigue testing. Two types of tests are performed: (1) fatigue testing of the strand and (2) testing of
relatively short lengths of the assembled cable with anchorages. The recommended test of the system is
2 million cycles at a stress range (158 MPa) that is 35 MPa greater than the fatigue allowable for
Category B at 2 million cycles. This test should pass with less than 2% wire breaks. A subsequent proof
test must achieve 95% of the guaranteed ultimate tensile strength of the tendons.

34.2.4 Low-Cycle Fatigue and Seismic Frame Details


Steel braced frames and moment-resisting frames are expected to withstand cyclic plastic deformation
without cracking in a large earthquake. If brittle fracture of these moment frame connections is suppressed
by good detailing and minimum notch toughness levels, as discussed in the next section, the connections
can be cyclically deformed into the plastic range and will eventually fail by tearing at a location of strain
concentration [65]. This limit state is referred to as low-cycle fatigue. Some examples of low-cycle fatigue
cracks are shown in Figure 34.6, Figure 34.7, Figure 34.8, and Figure 34.9 from full-scale beam-to-column

Copyright 2005 by CRC Press


34-12 Handbook of Structural Engineering

(a) (b)

FIGURE 34.6 Low-cycle fatigue crack developing at the toe of the beam flange weld in a moment–frame
connection after (a) 11 cycles and (b) 17 cycles of 4% drift.

LCF crack
Column Column
web flange

Slag inclusion,
LOF

FIGURE 34.7 Cross-section of beam flange weld showing low-cycle fatigue crack developing at the weld toe.

connection tests of the welded-unreinforced flange, welded web (WUF-W) connection, which is one of the
FEMA 350 prequalified connections for special moment frames in high seismic regions [66]. Figure 34.6
and Figure 34.7 show low-cycle fatigue cracks forming at the beam flange weld. Figure 34.8 shows a
low-cycle fatigue crack forming at the weld of the beam web to the column flange, and Figure 34.9 shows
a low-cycle fatigue crack forming at the weld access hole. The detailing in these connections is well balanced
since the low-cycle fatigue failure at all of these details occurred between 12 and 16 cycles of the maximum
required drift angle in each case.
Most past research on low-cycle fatigue has involved pressure vessels and some other types of
mechanical engineering structures. Since low-cycle fatigue is an inelastic phenomenon, the strain range
is the key parameter rather than the stress range. The Coffin–Manson rule [67] has been used to relate the
strain range in smooth tensile specimens to life. Manson suggested a conservative lower-bound simpli-
fication, called Manson’s universal slopes equation [68]:
su
De ¼ 3:5 N 0:12 þ ef0:6 N 0:6 ð34:5Þ
E
where De is the total strain range, su is the tensile strength, and ef is the elongation at fracture.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-13

FIGURE 34.8 Low-cycle fatigue crack developing at the end of the beam web to column flange weld in a moment–
frame connection.

FIGURE 34.9 Low-cycle fatigue crack developing at the weld access hole in a moment–frame connection.

Note that the first term in Equation 34.5 is the elastic part of the total strain range (which is relatively
insignificant when there are fewer than 100 cycles), and the second term is the plastic part of the total
strain range. Figure 34.10 shows a plot of Manson’s universal slopes equation where su is 450 MPa and
ef is 25%, typical minimum properties for Grade 50 structural steel. Many studies have shown that
Manson’s universal slopes equation is conservative compared to experimental data from smooth
specimens [68,69]. However, because of buckling at greater strain ranges, most of the experimental data
are for strain ranges less than 1%, that is, for cycles greater than 1000. Limited data exist at higher strain
ranges — some are shown in Figure 34.10 for A36 steel smooth specimens machined from the flanges of
wide-flange sections [69].

Copyright 2005 by CRC Press


34-14 Handbook of Structural Engineering

At present, very little is understood about low-cycle fatigue in welded or bolted structural details.
For example, it is a very difficult task just to predict accurately the local strain range at a location of cyclic
local flange buckling. However, Krawinkler and Zohrei [70] and Ballio and Castiglioni [71,72] showed
that the number of cycles to failure by low-cycle fatigue of welded connections could be predicted by the
local strain range in a power law that is analogous to an S–N curve. Ballio and Castiglioni [71,72] showed
that the power law would have an exponent of 3, just like the elastic S–N curves. Krawinkler and Zohrei
[70] also showed that Miner’s rule could be used to predict the number of variable-amplitude cycles to
failure based on constant-amplitude test data.
Therefore, it may be possible to predict and design against low-cycle fatigue using strain-range versus
number-of-cycles curves that are extrapolated from the high-cycle fatigue design S–N curves.
Figure 34.10 shows the AASHTO/AISC S–N curves for Categories A and C, converted from stress range
to strain range by dividing the stress ranges by the elastic modulus and extrapolated up to one cycle.
There are only limited data to support this approach. Figure 34.10 shows some data from the same
full-scale WUF-W beam-to-column connection tests for which the low-cycle fatigue cracks were shown
in Figure 34.6, Figure 34.7, Figure 34.8, and Figure 34.9. These tests are subjected to a standard series of
cycles of increasing ranges of total drift as shown in Figure 34.11. They are cycled at 4% drift until they

100
Manson’s equation Welded coupon
Smooth specimen
WUF-W connection tests
Category A
Strain range, %

10
Category C

0.1
1 10 100 1,000 10,000
Cycles

FIGURE 34.10 Comparison of standard S–N curves presented in terms of strain range and Manson’s universal
slopes equation for Grade 50 (350-MPa yield strength) steel to low-cycle fatigue test data.

0.05
Continue 4.0% drift
0.04 cycles to failure
0.03
Interstory drift angle, rad

0.02
0.01
0
–0.01
–0.02
–0.03
–0.04
–0.05
0 5 10 15 20 25 30
Cumulative cycles

FIGURE 34.11 Standard drift-cycle history for testing moment–frame subassemblies.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-15

fail by low-cycle fatigue. The measured flange strain ranges in these tests varied from 3.7 to 4.8% during
the 4% drift cycles. The number of cycles plotted in Figure 34.10 is the equivalent number of cycles at 4%
drift, which is the actual number of cycles at 4% plus one additional cycle which, according to Miner’s
rule, is equivalent in damage to all of the cycles at 3% and less. Since maximum flange strains were used
rather than nominal values, this is analogous to a hot-spot approach for high-cycle fatigue. As discussed
previously in Section 34.2.3.3, the Category C S–N curve is a suitable baseline S–N curve for the hot-spot
approach. It appears that the Category C S–N curve is also a good lower bound to these low-cycle fatigue
data. The scatter in the data is substantial, as is also true in high-cycle fatigue.
Also shown in Figure 34.10 are data for smaller coupon-type specimens with transverse butt welds,
which would be expected to be Category C details. These are some of the only available data with fewer
than 5000 cycles. These coupon data are also in reasonable agreement with the extrapolated Category C
curve as a lower bound.

34.3 Fracture
34.3.1 Fracture Resistance
Unlike fatigue, fracture behavior depends strongly on the type and strength level of the steel or filler
metal. In fact, the fracture resistance of each type of steel or weld metal varies significantly from heat to
heat, or from lot to lot. In fact, in rolled shapes, the fracture resistance may vary significantly within the
cross-section. Therefore, design or assessment for fracture resistance will usually involve a measurement
of the material’s fracture resistance, or fracture toughness. Although fracture toughness can be measured
directly in fracture mechanics tests, the usual practice is to characterize the toughness of steel in terms of
the impact energy absorbed by a Charpy V-notch (CVN) specimen. Because the Charpy test is relatively
easy to perform, it will likely continue to be the measure of toughness used in steel specifications.
Since it is not directly related to the fracture toughness, CVN energy is often referred to as notch
toughness. The notch toughness is still very useful, however, since it can often be correlated to the
fracture toughness and then used in a fracture mechanics assessment [3,32,73,74]. Figure 34.12 shows
a plot of the CVN energy of A588 Grade 50 (350-MPa yield strength) structural steel at varying
temperature. These results are typical for ordinary hot-rolled structural steel.
The fracture limit state includes phenomena ranging from brittle fracture of low-toughness materials
at service load levels to ductile tensile rupture of a component. The transition between these phenomena

Test temperature, °F
–150 –100 –50 0 50 100 150 200 250

90 A588 19-mm plate


60
Charpy data

70
CVN, ft-lb
CVN, J

40
50

30 20

10

–75 –50 –25 0 25 50 75 100


Test temperature, °C

FIGURE 34.12 Charpy transition curve for A588 Grade 50 (350-MPa yield strength) structural steel.

Copyright 2005 by CRC Press


34-16 Handbook of Structural Engineering

depends on temperature, as reflected by the variation of CVN with temperature as shown in Figure 34.12.
The transition is a result of changes in the underlying microstructural fracture mode.
Brittle fracture on the so-called lower shelf in Figure 34.12 is associated with cleavage of individual
grains on select crystallographic planes. Brittle fracture may be analyzed with linear-elastic fracture
mechanics theory because the plastic zone at the crack tip is very small.
At the high end of the temperature range, the so-called upper shelf, ductile fracture is associated with
the initiation, growth, and coalescence of microstructural voids, a process requiring much energy. The
net section of plates or shapes fully yields and then ruptures with large slanted shear lips on the fracture
surface.
Transition-range fracture occurs at temperatures between the lower and upper shelves and is asso-
ciated with a mixture of cleavage and shear fracture. Large variability in toughness at constant tem-
perature and large changes with temperature are typical of transition-range fractures.
Brittle fracture can be thought of as an interruption of what would otherwise be ductile plastic
deformation in tension, much like buckling can interrupt ductile plastic deformation in compression.
For the purposes of structural engineering, it is usually only necessary to make sure that brittle fracture
does not occur, so that is the emphasis in this chapter. As long as the behavior is in the transition range
or upper shelf, the resulting ductility is generally sufficient. For example, as discussed in the previous
section, moment-frame connections will tolerate a certain minimum number of cycles of large rotations
governed by the low-cycle fatigue limit state, as long as specific details are used and minimum CVN
requirements are assured. If low-toughness weld metal that does not meet minimum requirements is
used, brittle fracture may occur before the expected number of cycles [65].

34.3.2 Material Specification and Detailing to Avoid Fracture


Details that have good fatigue resistance, Category C and better, are usually also optimized for resistance
to fracture. Detailing rules to avoid fracture are very similar to the common-sense rules to avoid fatigue.
For example, intersecting welds should always be avoided due to the probability of defects and excessive
constraint. Intersecting welds, or even welds of too close proximity, have caused brittle fractures. For
example, in December 2000, fractures occurred in large plate girders of the Hoan Bridge in Milwaukee [75].
These fractures occurred at a shelf plate detail for attachment of the lateral wind bracing to the
girders where the shelf plate fit around, and intersected the welds of, a vertical stiffener, as shown in
Figure 34.13.
Fractures have also occurred at poorly executed weld access holes, as shown in Figure 34.14. AWS D1.1
[18] has detailing rules for base metal and welds that are good practice for bolted joints as well as welded
joints. Additional requirements are provided in the AISC Specifications [16,76] or other applicable
specifications. For example, AWS and AISC have specific rules for the size of weld access holes and the
radii and smoothness of all thermally cut edges.
The detailing rules of the AWS D1.1 and the AISC load and resistance factor design (LRFD) speci-
fication are sufficient for most statically loaded structures. More stringent detailing rules are also pro-
vided in AWS D1.5 [77] for bridges due to fatigue concerns. For example, weld backing bars must
usually be removed. AASHTO Construction Specifications [78] have some additional requirements,
for example, holes in primary load-carrying members must be drilled and not punched. This is
because punching may strain-harden the material (reducing toughness and ductility) and introduce
microcracks.
Also, more stringent detailing rules are provided in the AISC Seismic Specification [76] and in FEMA
350 [66] due to concerns about brittle fracture and low-cycle fatigue. For example, the connections
for special moment frames must be prequalified with specific details, such as the access hole visible in
Figure 34.8 and Figure 34.9.
In these and other applications requiring substantial ductility, stress concentrations such as re-entrant
corners should be avoided and instead transition radii that are ground smooth should be provided,
as shown in Figure 34.8 and Figure 34.9. The ends of butt welds are always a potential location of defects.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-17

FIGURE 34.13 Fractured girder of the Hoan Bridge in Milwaukee and view of critical shelf plate detail featuring
intersecting welds.

If possible, it is better to locate butt welds away from such stress concentrations. If not, it is important to
use runout tabs and to later grind the ends of the weld to a radius. Fillet weld terminations should not be
ground, for this will expose a very thin ligament near the weld root that will tear easily.
The enhanced detailing required by the bridge [77,78] or seismic [66,76] specifications are expensive
to implement relative to what would be required by AWS D1.1 [18] and AISC LRFD Specification [16].
Therefore, they should not be specified if extraordinary ductility is not required in a structural
component, which is the case for most statically loaded structures.
AASHTO specifications for bridge steel and weld filler metal and ASTM specifications for ship steel
(A131) require minimum CVN values at specific temperatures. However, since fatigue is not expected in
conventional buildings and therefore the risk of fracture is much lower, CVN is often not explicitly
specified for the steel or weld metal for buildings. Rather, the strategy is to allow only specific American
Society for Testing and Materials (ASTM) specifications for steel and AWS classifications for weld
filler metal that are known to function well in buildings. However, if there is special concern about
brittle fracture and (1) high-ductility demand (as in seismic frames), (2) low-temperature exposure,

Copyright 2005 by CRC Press


34-18 Handbook of Structural Engineering

FIGURE 34.14 Fracture emanating from a poorly executed weld access hole.

or (3) fatigue or impact loading, then a supplemental requirement for a minimum CVN at a specific
temperature should be considered, along with enhanced detailing. Since there is usually a premium cost,
a supplemental CVN requirement should only be specified when necessary.
As shown in Figure 34.12, the typical lower-shelf CVN is about 10 J. Therefore, when a minimum CVN
of 20 J or more is specified at some temperature, the most important result of such a specification is that
the lower shelf of the Charpy curve will start at a temperature lower than the specified temperature. In
fact, this indicates that the lower shelf of a structure loaded statically or at intermediate strain rates such
as traffic loading on a bridge is even lower, a phenomenon known as the temperature shift [3]. Because
of the temperature shift, the temperature at which the CVN requirement is specified may be greater than
the lowest anticipated service temperature.
As long as the material is not on the lower shelf at service temperature, brittle fracture will not occur as
long as large cracks do not develop. It almost does not matter what the specified CVN value is as long as
it is at least 20 J. Usually, an average from three tests of 34 J (25 ft lbs) or 27 J (20 ft lbs) is specified at a
particular temperature. The greater the value of the average CVN requirement, the more certain it is that
the material is well above the lower shelf, but there may be a greater premium to be paid with
diminishing increases in certainty.
In addition to brittle fracture, there is occasionally a problem with steel or weld metal that has low
upper-shelf toughness. Such a material can give a ductile failure mode but without sufficient ductility.
To guard against this type of material, a CVN requirement of 54 J (40 ft lbs) is often specified at 21
C
(room temperature).

34.3.2.1 Steel Plates and Shapes


AASHTO specifications for Grade 50 (minimum 350-MPa yield strength) bridge steel up to 38-mm
thick require a CVN at a temperature that is 38
C greater than the minimum service temperature. The
temperature shift accounts for the effect of strain rates, which are lower in the service loading of bridges
(on the order of 103) than in the Charpy test (greater than 101). The temperature shift means that when
a CVN is specified at some temperature, it ensures that the transition temperature for the structure is at
least 38
C less than the specified temperature. For bridge steel, the specified minimum CVN is 20 J for
nonfracture critical members and 34 J for fracture critical members. Fracture critical members are
defined as those that if fractured would result in collapse of the bridge.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-19

Most of the United States (except Alaska and a small part of the northern tier states) has a lowest
anticipated service temperature greater than 34
C, the limit for AASHTO Zone II. In Zone II, bridge
steel is required to have a minimum of 20 J at 4
C.
ASTM A673 has specifications for the frequency of Charpy testing. The H frequency requires a set of
three CVN specimens to be tested from one location for each heat or about 50 tons. These CVN test
specimens can be taken from a plate with thickness up to 9 mm different from the product thickness if it
is rolled from the same heat. The P frequency requires a set of three specimens to be tested from one end
of every plate, or from one shape in every 15 tons of that shape. For bridge steel, the AASHTO code
requires CVN tests at the H frequency as a minimum. For fracture critical members, CVN testing at the
P frequency is required.
For buildings, A36, A572, A588, or A992 do not have a specified minimum CVN, unless supplemental
specifications (CVN testing in accordance with ASTM A6/A6M, Supplementary Requirement S5) should
be cited.
One exception is shapes meeting A913, which are produced using a special quenched and self-
tempering process that results in good toughness. A913 shapes have a specified minimum CVN (average
of three tests) of 54 J at 21
C.
In most cases, ordinary structural steel has sufficient toughness for the required performance in
buildings, even under these demanding conditions. For example, Figure 34.15 shows a histogram of
CVN data at 4
C from the flanges of shapes for more than 2200 heats of A992 steel from five producers
in 1998 [79]. The CVN values were widely dispersed, and the mean values were typically very high, on the
order of 160 J or more. Table 34.2 shows some summary statistics for the A992 steel CVN. These data
show that 99.9% of the A992 steel meets Zone II, bridge steel requirement of 20 J at 4
C. Therefore, as
stated previously, it may not be worth paying the premium for specifying the supplemental CVN
requirement since the requirement is almost always met by all structural steel, at least that from the
major suppliers (U.S. and European) of the North American market.
Although the toughness of the flanges is generally quite good, the toughness of shapes is usually not
homogenous. For example, in rotary straightened W-shapes there is often an area of reduced notch
toughness in a limited region of the web immediately adjacent to the flange, referred to as the ‘‘k-area,’’
as illustrated in Figure 34.16 [80]. Following the 1994 Northridge earthquake, there was a tendency to
specify thicker continuity plates that were groove welded to the web and flange and thicker doubler
plates that were often groove welded in the gap between the doubler plate and the flanges. These welds
were highly restrained and may have caused cracking during fabrication in some cases [81].

50 100
Mean : 160
Minimum : 12
40 Maximum : 380 80
Coefficient of variation : 0.52
Cumulative probability, %

Number of samples : 2280


Occurrence, %

30 60

20 17 40
14
11 12 12
10 9
10 20
6 5
4
0.1
0 0
20 50 85 120 155 185 220 255 285 320 Over
CVN, J

FIGURE 34.15 CVN histogram for A992 steel at 4


C.

Copyright 2005 by CRC Press


34-20 Handbook of Structural Engineering

TABLE 34.2 Summary of Statistics for CVN for A992 Structural Steel
% less than 29 Joules First quartile (75% exceedence) Mean
Temperature (15 ft-lbs) Joules (ft-lbs) Joules (ft-lbs) Samples

0 C (32 F) 0.25 54 (40) 143 (106) 395


4
C (40
F) 0.13 88 (65) 163 (121) 2280
21
C (70
F) 0.09 104 (77) 175 (160) 1058

k-line

Area of
reduced
toughness

FIGURE 34.16 Location of the k-area affected by rotary straightening.

AISC issued an Advisory in 1997 [82] that recommended that the welds for continuity plates should
terminate well away from the k-area. The Advisory defined the k-area as the ‘‘region extending from
approximately the midpoint of the radius of the fillet into the web approximately 1 to 1.5 in beyond the
point of tangency between the fillet and web.’’
Recent pull-plate tests [83–86] and full-scale beam–column joint tests [65,86,87] have shown that this
problem can be avoided if the continuity plates are fillet welded to both the web and the flange, the
cutout or snipe at the corners of the continuity plates is at least 38 mm, and the fillet welds are stopped
short by a weld leg length from the edges of the cutout, as shown in Figure 34.17. These tests also show
that groove welding these continuity plates to the flanges or the web is unnecessary.
Tests have also shown the viability of fillet welding doubler plates to the flanges in lieu of groove welds
[65,84–86]. In most applications, if groove welds are not absolutely needed, fillet welds are almost always
preferred due to the lower restraint, lower residual stress, fewer defects, and associated reduced risk of
fracture.
Another relatively rare type of cracking in structural steel is galvanized cope cracking, which is thought
to be due to a phenomenon called liquid metal embrittlement. It has occurred at the edges of thermally
cut copes or rolled beams. The thermal cutting is believed to cause microcracking, allowing the molten
zinc and other metals to penetrate the microcracks and causing them to propagate into larger visible
cracks [88].
In the early 1980s, there were several fractures that occurred when jumbo sections with welded splices
were used as tension chords in long-span trusses [89]. Figure 34.18 shows an example of such a fracture
that originated at poorly cut access holes at the welded splices where they intersected the web/flange core
region of jumbo shapes [89]. This web/flange core region (not the same as the k-area) often had course

Copyright 2005 by CRC Press


Fatigue and Fracture 34-21

grain regions with very low toughness. Plates greater than 50 mm thick may also have regions of low
toughness at midthickness.
Therefore, AISC specifications now have special detailing requirements for weld access holes and
a supplemental Charpy requirement for shapes with flange thickness greater than 38 mm and plates
thicker than 51 mm, when these are welded and subject to primary tensile stress from axial load or
bending. These jumbo shapes and thick plates must exhibit an average of 27 J at 21
C. In the
shapes, the central longitudinal axis of the specimens must be located on a line in a plane one
fourth of the way through the thickness from the inside flange surface at the intersection with
the web midthickness. The test specimens are taken from the top of each ingot used to produce the
product.

5
16
in.
5
16
in.

5
16
in.

5
16
in.
1 12 in.

1 12 in.

FIGURE 34.17 Recommended placement of continuity plate fillet welds to avoid contact of welds with k-area.

FIGURE 34.18 Jumbo section used as tension chord in a roof truss and fracture in web originating from weld
access holes at welded splice.

Copyright 2005 by CRC Press


34-22 Handbook of Structural Engineering

34.3.2.2 Weld Filler Metal


Prior to the 1994 Northridge Earthquake, the welds in the welded special moment frame (WSMF)
connections were commonly made with the self-shielded flux-cored arc welding (FCAW-S) process
using an E70T-4 weld wire. Figure 34.19 shows a plot of CVN versus temperature for deposited
weld metal using E70T-4, E70TG-K2 (another FCAW-S filler metal), and ordinary E7018
shielded metal arc welding (SMAW) electrodes. The CVN for E70T-4 is less than 20 J for temperatures
up to 40
C [74].
Many of these low-toughness WSMF connections fractured in the earthquake [90]. Figure 34.20 shows
a fractured beam flange end and a cross-section. Figure 34.21 shows the crack-like notch created by the
backing bar. The fracture surfaces, such as those shown in Figure 34.20, indicate that the fractures
originate in the root of the weld, typically at a lack-of-fusion defect adjacent to the backing bar notch

Temperature, °C
–95 –55 –15 25 65 105
200
E70T-4 240
E70TG-K2
160 E7018
200
Absorbed energy, ft-lb

Absorbed energy, J
120 160

120
80
80
40
40

0 0
–140 –80 –20 40 100 160 220
Temperature, °F

FIGURE 34.19 Typical Charpy impact energy from E70T-4 FCAW-S weld metal from Northridge WSMF
connections compared to another FCAWS-S (E70TG-K2) and SMAW (E7018) weld metal.

FIGURE 34.20 Fracture surface and weld cross-section from moment–frame connection that fractured in the
Northridge Earthquake showing a typical crack that originated at the backing bar notch.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-23

Column
flange

CJP
weld

Primary
stress

Notch
perpendicular Girder
to primary flange
stress Backing
bar

FIGURE 34.21 Schematic cross-section of pre-Northridge groove weld detail with backing bar.

[91,92]. This lack-of-fusion defect is difficult to avoid when the weld must be stopped on one side of
the web and started on the other side.
A fracture mechanics assessment shows that the low-toughness E70T-4 weld metal would be predicted
to fracture in the presence of the crack-like notch at stress levels well below the yield point [73,74].
In addition to the low toughness and the crack-like notch and defect, there were many other important
factors in these fractures. The overall lack of redundancy, that is, the reliance on only one or two massive
WSMFs to resist lateral load in each direction, contributes to large forces, increase in the thickness of
the members, and the high constraint of the connections.
FEMA 350 has proposed improved connections, such as the WUF-W connection [66]. In addition to
specifying notch-tough filler metals, the detailing of the WUF-W connection is also improved relative to
the pre-Northridge connection. On the bottom flange weld, the backing bar must be removed, the weld
back gouged, and a reinforcing fillet placed underside. On the top flange, the backing bar may be left
in place, but the notch is sealed with a fillet weld as shown in Figure 34.7, reducing the adverse effect of
the notch.
The FEMA guidelines [66,93] recommend that weld metal be used that meets two CVN test
requirements: (1) 54 J at 21
C (to avoid low upper-shelf weld metal) and (2) 27 J at 18
C (to ensure
that the service temperature is above the lower shelf). Testing done as part of the FEMA program
suggests that another FCAW-S filler metal, E70T-6, can meet these requirements. The AWS classification
[94] requires E70T-6 to have 27 J at 29
C. The AISC seismic specifications [76] also require the weld
metal for these welds to have a minimum CVN of 27 J at 29
C. However, weld deposits with CVN
far less than required have been obtained under some conditions with E70T-6, despite certifications
indicating that the filler metal met the AWS classification requirements, leading to brittle fractures in
full-scale tests [65].
The FEMA requirements for minimum toughness are adequate, provided they can be consistently
met. Toughness is an inherently variable material property, particularly in a nonhomogenous material
such as a weld. For this reason, the toughness requirements should be treated as a lower-bound value
and not as an average. This can be accomplished either through strict quality assurance or specifi-
cation of welding consumables that have lower-bound toughness consistently above the FEMA
minimum.
The former is the approach required by the FEMA Recommended Specifications and Quality
Assurance Guidelines [93]. These Recommended Specifications require toughness testing on each

Copyright 2005 by CRC Press


34-24 Handbook of Structural Engineering

production lot of the specified filler metal. However, upon approval of the engineer, this requirement
may be waived and the consumable manufacturer’s certification testing may be used to verify the
material’s suitability [93].
The manufacturer’s testing may not be sufficient to preclude brittle weld deposits, since the testing
need only be conducted once a year on a single production lot of the particular electrode [94]. Alter-
natively, specification of higher toughness consumables (e.g., E71T-8) that consistently meet minimum
requirements may be a more reliable means of insuring welds of sufficient toughness.

34.4 Summary
1. Structural elements where the live load is a large percentage of the total load are potentially
susceptible to fatigue. Many factors in fabrication can increase the potential for fatigue, including
notches, misalignment and other geometrical discontinuities, thermal cutting, weld joint design
(particularly backing bars), residual stress, nondestructive evaluation and weld defects,
intersecting welds, and inadequate weld access holes.
2. The fatigue design procedures in the AASHTO and AISC specifications are based on control of
the stress range and knowledge of the fatigue strength of the various details. Using
these specifications, it is possible to identify and avoid details expected to have low fatigue
strength.
3. Low-cycle fatigue is a limit state for members and connections repeatedly cycled in the inelastic
range, such as for seismic loading. Low-cycle fatigue can be predicted using strain-range versus
cycles curves derived from the stress-based S–N curves for high-cycle fatigue.
4. Welded connections and thermal-cut holes copes, blocks, or cuts are potentially susceptible to
brittle fracture. Many interrelated design variables can increase the potential for brittle fracture,
including lack of redundancy, large forces and moments with dynamic loading rates, thick
members, geometrical discontinuities, and high constraint of the connections. Low temperature
can be a factor for exposed structures. The factors mentioned above, which influence the potential
for fatigue, have a similar effect on the potential for fracture. In addition, cold work (e.g., from
rotary straightening or punching holes), flame straightening, weld heat input, and weld sequence
can also affect the potential for fracture.
5. The AASHTO specifications require a minimum CVN notch toughness at a specified
temperature for the base metal and the weld metal of members loaded in tension or
tension due to bending. Almost two decades of experience with these bridge speci-
fications have proved that they are successful in significantly reducing the number of brittle
fractures.
6. Surveys of CVN for wide-flange shapes sold in North America show that 99.9% of this steel meets
the AASHTO bridge steel requirements for service down to  34
C. Therefore, under most
circumstances, there is no need to specify CVN for shapes used in buildings.
7. Achieving the required minimum CVN toughness in the girder flange-to-column flange groove
welds is critical for good performance in the prequalified steel moment connections. Lot testing
should be considered, or it may be necessary to specify electrodes that typically far exceed the
minimum toughness levels. Either step would assure that minimum CVN requirements represent
a realistic lower-bound of weld toughness deposited in the field.

References
[1] Bazant, Z. and Planas, J., Fracture and Size Effect in Concrete and Other Quasibrittle Materials, CRC
Press, Boca Raton, FL, 1997.
[2] Anderson, T.L., Fracture Mechanics — Fundamentals and Applications, Second Edition, CRC
Press, Boca Raton, FL, 1995.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-25

[3] Barsom, J.M. and Rolfe, S.T., Fracture and Fatigue Control in Structures: Applications of Fracture
Mechanics, Third Edition, ASTM, West Conshohocken, PA, 1999.
[4] Broek, D., Elementary Fracture Mechanics, Fourth Edition, Martinis Nijhoff Publishers, Dordrecht,
Netherlands, 1987.
[5] Kober, G.R., Dexter, R.J., Kaufmann, E.J., Yen, B.T., and Fisher, J.W., ‘‘The Effect of Welding
Discontinuities on the Variability of Fatigue Life,’’ Fracture Mechanics, Twenty-Fifth Volume,
ASTM STP 1220, F. Erdogan and R.J. Hartranft, Eds., American Society for Testing and Materials,
Philadelphia, PA, 1994.
[6] Fisher, J.W., Frank, K.H., Hirt, M.A., and McNamee, B.M., Effect of Weldments on The Fatigue
Strength of Steel Beams, National Cooperative Highway Research Program (NCHRP) Report 102,
Highway Research Board, Washington, DC, 1970.
[7] Fisher, J.W., Albrecht, P.A., Yen, B.T., Klingerman, D.J., and McNamee, B.M., Fatigue Strength of
Steel Beams with Welded Stiffeners and Attachments, National Cooperative Highway Research
Program (NCHRP) Report 147, Transportation Research Board, Washington, DC, 1974.
[8] Keating, P.B. and Fisher, J.W., Evaluation of Fatigue Tests and Design Criteria on Welded Details,
National Cooperative Highway Research Program (NCHRP) Report 286, Transportation
Research Board, Washington, DC, 1986.
[9] Maddox, S.J., Fatigue Strength of Welded Structures, Second Edition, Abington Publishing,
Cambridge, U.K., 1991.
[10] Dexter, R.J., Fisher, J.W. and Beach, J.E., ‘‘Fatigue Behavior of Welded HSLA-80 Members,’’
Proceedings, 12th International Conference on Offshore Mechanics and Arctic Engineering, Vol. III,
Part A, Materials Engineering, M.M. Salama et al., Eds., pp. 493–502, ASME, New York, 1993.
[11] Petershagen, H. and Zwick, W., Fatigue Strength of Butt Welds Made by Different Welding Processes,
IIW-Document XIII-1048-82, 1982.
[12] Petershagen, H., ‘‘The Influence of Undercut on the Fatigue Strength of Welds — A Literature
Survey,’’ Weld. World, Vol. 28, No. 7/8, pp. 29–36, 1990.
[13] Metrovich, B., Fisher, J.W., Yen, B.T., Kaufmann, E.J., Cheng, X., and Ma, Z., ‘‘Fatigue Strength of
Welded AL-6XN Superaustenitic Stainless Steel,’’ Int. J. Fatigue, Vol. 25, No. 9–11, pp. 1309–1315.
[14] Kaczinski, M.R., Dexter, R.J., and Van Dien, J.P., Fatigue-Resistant Design of Cantilevered Signal,
Sign, and Light Supports, National Cooperative Highway Research Program, NCHRP Report 412,
Transportation Research Board, Washington, DC, 1998.
[15] AASHTO, AASHTO LRFD Bridge Design Specifications, Second Edition, The American Association
of State Highway and Transportation Officials, Washington, DC, 1998.
[16] AISC, Load and Resistance Factor Design Specification for Structural Steel Buildings, Third Edition,
American Institute of Steel Construction (AISC), Chicago, 1999.
[17] AREMA, AREMA Manual for Railway Engineering, Chapter 15: ‘‘Steel Structures,’’ American
Railway Engineering and Maintenance of Way Association, 2002.
[18] AWS, Structural Welding Code — Steel, ANSI/AWS D1.1-02, American Welding Society, Miami,
FL, 2002.
[19] CSA, CSA S16-2001, Limit States Design of Steel Structures, Canadian Standards Association,
Toronto, Ontario, 2001.
[20] CEN, ENV 1999-1-1, Eurocode 3: Design of Steel structures — Part 1.1: General Rules and Rules for
Buildings, European Committee for Standardization (CEN), Brussels, May 1998.
[21] BSI, BS 7608, Specification of Practice for Fatigue Design and Assessment of Steel Structures, British
Standards Institute, London, 1994.
[22] ABS, Guide for Fatigue Strength Assessment of Tankers, American Bureau of Shipping, New York,
June 1992.
[23] UK Health & Safety Executive (formerly the UK Department of Energy), Fatigue Design Guidance
for Steel Welded Joints in Offshore Structures, H.M.S.O., London, 1984.
[24] Roberts, R. et al., Corrosion Fatigue of Bridge Steels, Vols. 1–3, Reports FHWA/RD-86/165, 166,
and 167, Federal Highway Administration, Washington, DC, May 1986.

Copyright 2005 by CRC Press


34-26 Handbook of Structural Engineering

[25] Outt, J.M.M., Fisher, J.W., and Yen, B.T., Fatigue Strength of Weathered and Deteriorated Riveted
Members, Report DOT/OST/P-34/85/016, Department of Transportation, Federal Highway
Administration, Washington, DC, October 1984.
[26] Albrecht, P. and Shabshab, C., ‘‘Fatigue Strength of Weathered Rolled Beam Made of A588 Steel,’’
J. Mater. Civ. Eng., Vol. 6, No. 3, pp. 407–428, 1994.
[27] The Aluminum Association, The Aluminum Design Manual, The Aluminum Association,
Washington, DC, 1986.
[28] Menzemer, C.C. and Fisher, J.W., ‘‘Revisions to the Aluminum Association Fatigue Design
Specifications, 6th International Conference on Aluminum Weldments, Cleveland, OH, April 3–5,
1995, AWS, Miami, FL, pp. 11–23, 1995.
[29] Fisher, J.W., et al., Resistance of Welded Details under Variable Amplitude Long-Life Fatigue
Loading, National Cooperative Highway Research Program Report 354, Transportation Research
Board, Washington, DC, 1993.
[30] Dexter, R.J., Wright, W.J., and Fisher, J.W., ‘‘Fatigue and Fracture of Steel Girders,’’ J. Bridge Eng.,
Vol. 9, Issue 3, May/June 2004, pp. 278–286.
[31] Miner, M.A., ‘‘Cumulative Damage in Fatigue,’’ J. Appl. Mech., Vol. 12, A-159, 1945.
[32] BS 7910, Guide on Methods for Assessing the Acceptability of Flaws in Metallic Structures, British
Standards Institute, London, 1999.
[33] Demers, C. and Fisher, J.W., Fatigue Cracking of Steel Bridge Structures, Volume I: A Survey
of Localized Cracking in Steel Bridges – 1981 to 1988, Report No. FHWA-RD-89-166, and Volume II:
A Commentary and Guide for Design, Evaluation, and Investigating Cracking, Report No. FHWA-
RD-89-167, FHWA, McLean, VA, March 1990.
[34] Yen, B.T., Huang, T., Lai, L.-Y., and Fisher, J.W., Manual for Inspecting Bridges for Fatigue Damage
Conditions, Report No. FHWA-PA-89-022 þ 85-02, Fritz Engineering Laboratory Report
No. 511.1, Pennsylvania Department of Transportation, Harrisburg, PA, January 1990.
[35] Dexter, R.J. and Kelly, B.A., ‘‘Research on Repair and Improvement Methods,’’ International
Conference on Performance of Dynamically Loaded Welded Structures, Proceedings of the IIW 50th
Annual Assembly Conference, San Francisco, CA, July 13–19, 1997, Welding Research Council,
Inc., New York, pp. 273–285, 1997.
[36] Kelly, B.A. and Dexter, R.J., Adequacy of Weld Repairs for Ships, SSC-424, Ship Structure
Committee, Washington, DC, 2003.
[37] Frank, K.H. and Fisher, J.W., ‘‘Fatigue Strength of Fillet Welded Cruciform Joints,’’ J. Struct. Div.,
ASCE, Vol. 105, No. ST9, pp. 1727–1740, September 1979.
[38] Kulak, G.L., Fisher, J.W., and Struick, J.H., Guide to Design Criteria for Bolted and Riveted Joints,
Second Edition, Prentice Hall, Englewood Cliffs, NJ, 1987.
[39] Kuperus, A., The Fatigue Strength of Tensile Loaded Non-Tightened HSFG Bolts. Delft University of
Technology, Report 6-73-3, June 1973.
[40] Kuperus, A., The Fatigue Strength of Tensile Loaded Tightened HSFG Bolts. Delft University of
Technology, Report 6-74-4, October 1974.
[41] Frank, K.H., ‘‘Fatigue Strength of Anchor Bolts,’’ J. Struct. Div., ASCE, Vol. 106, No. ST, June 1980.
[42] Dusel, J.P., Stoker, J.P. and Travis, R., Determination of Fatigue Characteristics of Hot-Dipped
Galvanized A307 and A449 Anchor Bars and A325 Cap Screws, California Department of Trans-
portation Reports, February 1984.
[43] Van Wingerde, A., Packer, J., and Wardenier, J., ‘‘Criteria for the Fatigue Assessment of Hollow
Structural Section Connections,’’ J. Constr. Steel Res., Vol. 35, pp. 71–115, 1995.
[44] Van Wingerde, A., Packer, J., and Wardenier, J., ‘‘New Guidelines for Fatigue Design of HSS
Connections,’’ J. Struct. Eng., Vol. 122, No. 2, pp. 125–132, February 1996.
[45] Zhao, X. and Packer, J., Fatigue Design Procedure for Welded Hollow Section Joints, International
Institute of Welding documents XIII-1804-99 and IIW Document XV-1035-99, Abington
Publishing, Cambridge, U.K., 2000.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-27

[46] Zhao, X. et al., Design Guide for Circular and Rectangular Hollow Section Welded Joints under Fatigue
Loading, Comite International pour le Developpement et l’Etude de la Construction Tubulaire, 1999.
[47] AASHTO, Standard Specifications for Structural Supports for Highway Signs, Luminaires and Traffic
Signals, Fourth Edition, American Association of State Highway and Transportation Officials,
Washington, DC, 2001.
[48] Marshall, P., ‘‘Welded Tubular Connections — CHS Trusses,’’ Handbook of Structural Engineering,
W.F. Chen, Ed., CRC Press, Boca Raton, FL, 1999.
[49] Dexter, R., Tarquinio, J., and Fisher, J., ‘‘Application of Hot-Spot Stress Fatigue Analysis to
Attachments on Flexible Plate,’’ Proceedings of the 13th International Conference on Offshore
Mechanics and Arctic Engineering Conference (OMAE), M.M. Salama et al., Eds., American Society
of Mechanical Engineers, Materials Engineering, Vol. III, pp. 85–92, 1994.
[50] Yagi, J., Machida, S., Tomita, Y., Matoba, M., Kawasaki, T., Definition of Hot-Spot Stress in
Welded Plate Type Structure for Fatigue Assessment, International Institute of Welding, IIW-XIII-
1414-91, 1991.
[51] ACI Committee 215, ‘‘Considerations for Design of Concrete Structures Subjected to Fatigue
Loading,’’ ACI 215R-74 (Revised 1992), ACI Manual of Standard Practice, Part 1, 2000.
[52] Hahin, C., Effects of Corrosion and Fatigue on the Load-Carrying Capacity of Structural Steel and
Reinforcing Steel, Illinois Physical Research Report No. 108, Illinois Department of Transporta-
tion, Springfield, IL, March 1994.
[53] Pfister, J.F. and Hognestad, E., ‘‘High Strength Bars as Concrete Reinforcement, Part 6, Fatigue
Tests,’’ J. PCA Res. Dev. Lab., Vol. 6, No. 1, pp. 65–84, January 1964.
[54] Sternberg, F., Performance of Continuously Reinforced Concrete Pavement, I-84 Southington,
Connecticut State Highway Department, June 1969.
[55] Batchelor, B., Hewitt B.E., and Csagoly, P., ‘‘An Investigation of the Fatigue Strength of Deck Slabs
of Composite Steel/Concrete Bridges,’’ Transportation Research Record 664, Vol. 1, TRB, National
Research Council, Washington, DC, pp. 153–161, 1978.
[56] Azad, A.K., Baluch M.H., Al-Mandil M.Y., and Al-Suwaiyan, ‘‘Static and Fatigue Tests of
Simulated Bridge Decks,’’ Experimental Assessment of Performance of Bridges, L.R. Wang and
G.M. Salnis, Eds., Proceedings of ASCE Convention, Boston, MA, pp. 30–41, October 1986.
[57] Youn, S. and Chang S., ‘‘Behavior of Composite Bridge Decks Subjected to Static and Fatigue
Loading,’’ ACI Struct. J., Vol. 95, No. 3, pp. 249–259, 1998.
[58] Okada, K., Okamura H., and Sonada K., ‘‘Fatigue Failure Mechanism of Reinforced Concrete
Deck Slabs,’’ Transportation Research Record 664, National Research Council, Washington, DC,
pp. 136–144, 1978.
[59] Kato, T. and Goto Y., ‘‘Effect of Water Infiltration of Penetration Deterioration of Bridge
Deck Slabs,’’ Transportation Research Record 950, National Research Council, Washington, DC,
pp. 202–209, 1978.
[60] Petrou, M., Perdikaris P.C., and Wang, A., ‘‘Fatigue Behavior of Non-Composite Reinforced
Concrete Bridge Deck Models,’’ Transportation Research Record 1460, National Research Council,
Washington, DC, pp. 73–80, 1994.
[61] Fang, I.K., Tsui, C.K.T., Burns N.H., and Klinger, R.E., ‘‘Fatigue Behavior of Cast-in-Place
and Precast Panel Bridge Decks with Isotropic Reinforcement,’’ PCI J., Vol. 35, No. 3, pp. 28–39,
May–June 1990.
[62] Rabbat, B.G., et al., ‘‘Fatigue Tests of Pretensioned Girders with Blanketed and Draped Strands,’’
J., Prestressed Concrete Inst., Vol. 24, No. 4, pp. 88–115, July/August 1979.
[63] Overnman, T.R., Breen, J.E., and Frank, K.H., Fatigue Behavior of Pretensioned Concrete Girders,
Research Report 300-2F, Center for Transportation Research, The University of Texas at Austin,
November 1984.
[64] Ad-Hoc Committee on Cable-Stayed Bridges, Recommendations for Stay Cable Design, Testing,
and Installation. Fourth Edition, Post-Tensioning Institute, Phoenix, AZ, March 2001.

Copyright 2005 by CRC Press


34-28 Handbook of Structural Engineering

[65] Lee, D., Cotton, S., Dexter, R.J., Hajjar, J.F., Ye, Y., and Ojard, S.D., Column Stiffener Detailing and
Panel Zone Behavior of Steel Moment Frame Connections, Report No. ST-01-3.2, Department of
Civil Engineering, University of Minnesota, Minneapolis, MN, 2002.
[66] FEMA, Recommended Seismic Design Criteria for New Steel Moment-Frame Buildings, Report
No. FEMA 350, Federal Emergency Management Agency, Washington, DC, 2000.
[67] Coffin, L.F., Jr., ‘‘A Note on Low Cycle Fatigue Laws,’’ J. Mater., Vol. 6, No. 2, pp. 388–402, 1971.
[68] Itoh, Y.Z. and Kashiwaya, H., ‘‘Low-Cycle Fatigue Properties of Steel Sand Their Weld Metals,’’
J. Eng. Mater. Technol., Vol. 111, No. 4, pp. 431–437, 1989.
[69] Howdyshell, P., Trovillion, J.C., and Wetterich, J.L., ‘‘Low-Cycle Fatigue of Structural Materials,’’
Materials and Construction: Proceedings of MatCong 5, The 5th ASCE Materials Engineering
Congress, Cincinnati, OH, 10–12 May 1999, L. Bank, Ed., American Society of Civil Engineers,
Reston, VA, pp. 148–155.
[70] Krawinkler, H. and Zohrei, M., ‘‘Cumulative Damage in Steel Structures Subjected to Earthquake
Ground Motion,’’ Comput. Struct., Vol. 16, No. 1–4, pp. 531–541, 1983.
[71] Ballio, G. and Castiglioni, C.A., ‘‘A Unified Approach for the Design of Steel Structures under Low
and/or High Cycle Fatigue’’, J. Constr. Steel Res., Vol. 34, No. 1, pp. 75–101, 1995.
[72] Castiglioni, C.A., ‘‘Cumulative Damage Assessment in Structural Steel Details,’’ IABSE Sympo-
sium, San Francisco, 1995, Extending the Lifespan of Structures, pp. 1061–1066, 1995.
[73] Dexter, R.J. and J.W. Fisher, ‘‘Fatigue and Fracture,’’ Chapter 8, Steel Design Handbook, LRFD
Method, A.R. Tamboli, Ed., McGraw-Hill, New York, 1997.
[74] Fisher, J.W., Dexter, R.J., and Kaufmann, E.J., Fracture Mechanics of Welded Structural Steel
Connections, Report No. SAC 95-09, FEMA-288, March 1997.
[75] Wright, W.J., Fisher, J.W., and Sivakumar, B., Hoan Bridge Failure Investigation, Federal Highway
Administration, Washington, DC, 2001.
[76] AISC, Seismic Provisions for Structural Steel Buildings, Second Edition, American Institute of Steel
Construction (AISC), Chicago, 2002.
[77] AWS, Bridge Welding Code, AASHTO-AWS-D1.5M-D1.5: 2002, American Welding Society,
Miami, FL, 2002.
[78] AASHTO, LRFD Bridge Construction Specifications, First Edition, The American Association of
State Highway and Transportation Officials, Washington, DC, 1998.
[79] Dexter, R.J., Structural Shape Material Property Survey, Final Report to Structural Shape Produ-
cer’s Council, University of Minnesota, Minneapolis, MN, 2000.
[80] Kaufmann, E.J., Metrovich, B., Pense, A.W., and Fisher, J.W., ‘‘Effect of Manufacturing Process on
k-Area Properties and Service Performance,’’ Proceedings of the North American Steel Construction
Conference, Fort Lauderdale, FL, May 9–12, 2001, AISC, Chicago, IL, pp. 17-1–17-24, 2001.
[81] Tide, R.H., ‘‘Evaluation of Steel Properties and Cracking in the ‘k’-Area of W Shapes,’’ Eng. Struct.,
Vol. 22, pp. 128–124, 1999.
[82] AISC, ‘‘AISC Advisory Statement on Mechanical Properties near the Fillet of Wide Flange Shapes
and Interim Recommendations January 10, 1997,’’ Modern Steel Construction, p. 18, February 1997.
[83] Dexter, R.J. and Melendrez, M.I., ‘‘Through-Thickness Properties of Column Flanges in Welded
Moment Connections,’’ J. Struct. Eng., ASCE, Vol. 126, No. 1, pp. 24–31, 2000.
[84] Prochnow, S.D., Ye, Y., Dexter, R.J., Hajjar, J.F., and Cotton, S.C., ‘‘Local Flange Bending and
Local Web Yielding Limit States in Steel Moment Resisting Connections,’’ Connections in Steel
Structures IV, Roanoke, VA, October 22–25, 2000, R.T. Leon and W.S. Easterling, Eds., AISC,
Chicago, 2002.
[85] Hajjar, J.F., Dexter, R.J., Ojard, S.D., Ye, Y., and Cotton, S.C., ‘‘Continuity Plate Detailing for Steel
Moment-Resisting Connections,’’ Eng. J., Vol. 40, No. 4, 4th Qrt., pp. 189–211, 2003.
[86] Dexter, R.J., Hajjar, J.F., Prochnow, S.D., Graeser, M.D., Galambos, T.V., and Cotton, S.C.,
‘‘Evaluation of the Design Requirements for Column Stiffeners and Doublers and the Variation in
Properties of A992 Shapes,’’ 2001 North American Steel Construction Conference, Ft. Lauderdale,
FL, May 9–12, 2001, AISC, Chicago, 2001.

Copyright 2005 by CRC Press


Fatigue and Fracture 34-29

[87] Bjorhovde, R., Goland, L.J., and Benac, D.J., ‘‘Performance of Steel in High-Demand Full-Scale
Connection Tests,’’ 2000 North American Steel Construction Conference, Las Vegas, NV, May 2000,
AISC, pp. 3.1–3.22, February 2000.
[88] Langill, T.J. and Schlafly, T., ‘‘Cope Cracking in Structural Steel after Galvanizing,’’ Mod. Steel
Constr. (USA), Vol. 35, No. 10, pp. 40–43, October 1995.
[89] Fisher, J.W. and Pense, A.W., ‘‘Experience with Use of Heavy W Shapes in Tension,’’ Eng. J., AISC,
Vol. 24, No. 2, 1987.
[90] Northridge Reconnaissance Team (1996). Northridge Earthquake of January 17, 1994, Recon-
naissance Report (Supplement C-2 to Volume 11), EERI, Oakland, CA.
[91] Kaufmann, E.J., Fisher, J.W., Di Julio, R.M., Jr., and Gross, J.L., Failure Analysis of Welded Steel
Moment Frames Damaged in the Northridge Earthquake, NISTIR 5944, National Institute of
Standards and Technology, Gaithersburg, MD, January 1997.
[92] Tide, R.H.R., Fisher, J.W., and Kaufmann, E.J., ‘‘Substandard Welding Quality Exposed:
Northridge, California Earthquake, January 17, 1994,’’ IIW Asian Pacific Welding Congress,
Auckland, New Zealand, February 4–9, 1996.
[93] FEMA, Recommended Specifications and Quality Assurance Guidelines for Steel Moment-Frame
Construction for Seismic Applications, Report No. FEMA 353, FEMA, Washington, DC, 2000.
[94] AWS, AWS A5.20–95, Specification for Carbon Steel Electrodes for Flux Cored Arc Welding,
American Welding Society AWS, Miami, FL, 1995.

Copyright 2005 by CRC Press


Copyright 2005 by CRC Press

You might also like