You are on page 1of 15

60th Anniversary State-of-the-Art Reviews

Predicting Fracture in Civil Engineering


Steel Structures: State of the Art
Amit Kanvinde, M.ASCE 1

Abstract: Fracture is an extreme limit state in steel structures, often precipitating structural failure or serious loss of function. Methods
to predict fracture in civil structures include traditional approaches developed in other disciplines (mechanical or aerospace engineering)
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

subsequently adopted in structural engineering, as well as approaches to characterize earthquake-induced fractures originating within civil
engineering. Developed over nearly six decades, the state of the art is composed of theories and models that address fracture over multiple
scales and are targeted toward disparate application scenarios. The paper examines these approaches from a structural engineering standpoint,
considering trade-offs in accuracy and expense, while identifying areas for improvement. Traditional approaches (including linear elastic and
elastic plastic fracture mechanics) are presented, followed by newer local approaches that are better suited for scenarios where traditional
approaches are inapplicable. By simulating micromechanisms such as microvoid growth as well as granular cleavage, local approaches
address fracture under large-scale yielding, ultralow-cycle fatigue (which occurs during earthquakes), and low-stress triaxiality, all of which
are important in civil structures. The physical basis for these approaches is outlined, with a summary of best practices for calibration and
application. However, these local approaches have limitations as well, and often require substantial resources for successful implementation.
With this background, optimal fracture assessment strategies are outlined for common structural scenarios, considering accuracy and cost.
Limitations of the entire fracture modeling framework are summarized because they pertain to mainstream adoption within structural
engineering research and practice. As the profession moves toward accurate performance characterization, it is anticipated that research
will accelerate to overcome these limitations. DOI: 10.1061/(ASCE)ST.1943-541X.0001704. © 2016 American Society of Civil Engineers.
Author keywords: Fracture; Steel structures; Metal and composite structures.

Introduction (structural) engineering, e.g., earthquake-induced ultralow-cycle


fatigue (ULCF). The subsequent evolution of fracture mechanics
Failure of steel structures triggered by fracture is a much-feared within structural engineering has been influenced by various factors
situation across various engineering disciplines. Beginning with the including (1) its application scenario, e.g., sharp versus blunt
Liberty Ship fractures in the 1940s (Sheldon 2010) that catalyzed cracked components, monotonic versus cyclic loading, and limited
the initial development of fracture mechanics as a field of study, versus extensive yielding within the component; (2) computational
catastrophic structural failures are periodic reminders of the conse- resources available at the time of its development; and (3) estab-
quences of fracture. Prominent examples include the fracture of the lished trade practices and standards for material testing and char-
Alexander Kielland drilling rig in 1980 (Almar-Naess et al. 1984) acterization. Not surprisingly, fracture mechanics sometimes
resulting in 123 deaths, and steel building fractures during the 1994 appears to be a patchwork of competing theories and methodolo-
Northridge and 1995 Kobe earthquakes. Because some of these gies whose suitability for addressing structural fracture under vari-
earlier fractures occurred outside the realm of civil infrastructure, ous conditions is somewhat unclear. In view of this, the primary
fracture mechanics initially evolved within the mechanical, aero- objective of this paper is to present fracture prediction approaches
space, and nuclear engineering disciplines. Within civil engineer- from a structural engineering standpoint with consideration of the
ing, fracture was largely addressed through prescriptive detailing following factors: (1) application scenario, (2) trade-offs between
or through design checks in which applied stresses (calculated expense and accuracy, (3) limitations and areas for refinement, and
through classical strength of materials–type approaches) were (4) improvements to professional practices to enable more effective
checked against empirical stress capacities (American Welding fracture simulation.
Society 2015). The Northridge and Kobe earthquakes underscored The next section provides a general overview of fracture in steel
the problems with these empirical approaches, providing impetus structures, describing the responsible micromechanisms across
for studying fracture within a context specific to civil (especially multiple scales. The subsequent section outlines methodologies for
earthquake) engineering. This resulted in adaptation of existing fracture prediction, recognizing that all ultimately seek to charac-
methodologies to civil engineering (Barsom and Rolfe 1999), as terize common physical mechanisms, albeit in varying contexts and
well as the development of new models and methods (Kanvinde at different scales, such that these methods are suited for different
and Deierlein 2007) that address problems peculiar to civil application scenarios. The next section examines scenarios (mate-
1
rials, loading types, geometry) common to structural engineering
Professor, Dept. of Civil and Environmental Engineering, Univ. of and the suitability of the different fracture models to address them.
California, Davis, CA 95616. E-mail: kanvinde@ucdavis.edu
The paper then summarizes limitations of the state of the art as they
Note. This manuscript was submitted on February 4, 2016; approved
on September 21, 2016; published online on November 11, 2016. pertain to structural engineering, outlining areas for improvement
Discussion period open until April 11, 2017; separate discussions must in the modeling methods themselves and in professional practices
be submitted for individual papers. This paper is part of the Journal of that are necessary to fully leverage these methods. Finally, notewor-
Structural Engineering, © ASCE, ISSN 0733-9445 thy applications of fracture mechanics are discussed to illustrate its

© ASCE 03116001-1 J. Struct. Eng.

J. Struct. Eng., 03116001


growing acceptance in professional practice. The focus of the paper such as the stress intensity factor (Anderson 1995) or the J-integral
is to review the state of the art for predicting fracture in civil steel (Rice 1968) are calculated from crack tip stress fields and used as
structures, and not the state of the art of fracture mechanics itself. indicators of fracture toughness demands to be compared with their
This is an important distinction because the former is subject to counterpart capacities, which are determined from laboratory tests.
subjectivities in material characterization owing to trade practices, At the other end, local fracture criteria are represented as critical
as well as a relative lack of resources (compared to other disciplines combinations of stress and/or strain at the location of crack initia-
and applications) for effective application of some of the most ad- tion (rather than from the remote stress-strain field). In contrast to
vanced fracture mechanics approaches. Consequently, the paper the higher-scale models, the lower-scale models can be generalized
emphasizes approaches that are practical for structural engineering across a wider range of geometrical configurations, material prop-
while discussing the advanced models to provide fuller context. erties, and loading types (e.g., monotonic versus cyclic). However,
they require simulation of stress-strain fields at a high resolution,
entailing high computational cost. Moreover, they require depth of
Fracture in Steel Structures: General Overview expertise in continuum finite-element simulation, large deforma-
tion mechanics, and constitutive modeling. Given these trade-offs,
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

From a structural engineering standpoint, fracture is interpreted as the optimal model depends on the structural configuration, type of
the separation of material leading to loss in load-carrying capacity. loading, and desired level of accuracy relative to resources available
The goal of fracture mechanics is to predict this event in terms of for material testing and simulation. This paper provides an assess-
applied loads or deformations. The fracture event itself may be de- ment of various methodologies against this backdrop, considering
composed into phenomena leading up to it: (1) concentration of (1) their fidelity to the physics of fracture, (2) their accuracy and
stress and strain fields over small regions of the component due generality across various scenarios, and (3) their cost and practical-
to stress increasers such as cracks or notches, or other phenomena ity. The next section describes the physical phenomena responsible
such as local buckling or necking; and (2) subsequent activation of for fracture at the microstructural scale and their correspondence
microstructural phenomena such as microvoid nucleation, growth, with macrostructural response.
and coalescence, or alternatively granular cleavage within the local-
ized zone, to initiate and propagate fracture. Simulation of the for-
mer (development of strain-stress fields and their concentration) is Mechanisms and Physics of Fracture in
typically within the scope of continuum finite-element simulation Structural Steel
with appropriate sophistication (large deformations, multiaxial
plasticity). On the other hand, the latter (microstructural phenom- Referring to the previous section, fracture criteria seek to establish
ena resulting in material separation) must be inserted into the analy- conditions (at the component or continuum scale) under which
sis as failure or damage criteria. This may be performed in an microstructural fracture mechanisms are activated. This section
uncoupled sense, in which the failure criterion is applied after the provides an overview of these mechanisms. When the steel micro-
stress-strain fields have been computed, with the presumption that structure (consisting of metal grains along with carbide and sulfide
once fracture initiates propagation and failure are inevitable. Alter- inclusions) is subjected to sufficient stress or strain, material sep-
natively, this may be done in a coupled manner, where material aration occurs through one of two processes or their combination:
separation is explicitly simulated to capture its interactions with the (1) cleavage due to tensile stress rupturing molecular bonds within
surrounding stress and strain fields. Typically, this is done through or at boundaries of grains, and (2) microvoid growth and coales-
one of the following constructs: (1) creation of free surfaces in cence due to inelastic flow of material around initiated microvoids,
the simulation model (Baldwin and Rashid 2013), (2) a softening leading to their enlargement and subsequent coalescence to form
constitutive model (e.g., Gurson 1977) along with finite-element the fracture surface. Figs. 1(a and b) schematically illustrate these
removal, or (3) a cohesive zone model (Elices et al. 2002). These processes, as well as the resultant fracture surfaces. Cleavage frac-
approaches enable the simulation of ductile crack growth and its ture does not require significant inelastic dissipation to propagate,
transition to unstable, brittle propagation (Ruggieri et al. 1996). whereas microvoid growth and coalescence is accompanied by
The section on crack propagation presented subsequently provides plastic dissipation. The implication of this at the macrostructural
further discussion of these methods. The fracture criterion itself is scale is that cleavage often propagates in an unstable manner (with-
usually expressed as a comparison between toughness capacity out increase in applied loads), whereas microvoid controlled frac-
(which is assumed a material property and may be defined in multi- ture, requiring higher energy, propagates slowly and stably with
ple ways) and a corresponding fracture toughness demand. The respect to applied loads. As a result, cleavage is often referred to
former is determined through laboratory tests, whereas the latter as brittle, whereas tearing is often considered ductile. While true
(which is a function of applied loads and geometry) is determined in general, the suddenness of fracture at the macroscale, often per-
through computational simulation, or, for more regular geometries ceived as brittleness, is not necessarily associated with cleavage
and materials, through closed-form equations. The multiplicity of and is controlled by factors such as strain energy stored in the en-
fracture criteria arises from the scales at which they are applied and tire component, loading rate, and its interaction with the energy
the mechanisms they seek to simulate. required to propagate the crack locally. For example, under high
Fracture criteria that are expressed at the higher (macro) scale loading rates or load-controlled conditions, even a locally ductile
cannot be generalized easily across various geometric configura- fracture mechanism (microvoid growth) can result in highly non-
tions or loading types. These include experimental methods, in ductile performance (sudden fracture) at the structural scale.
which laboratory measurements of load or deformation correspond- The preferred process depends on three factors: temperature,
ing to fracture are used as direct estimates of the fracture capacity of loading rate, and stress constraint (stress triaxiality). Lower temper-
the structural component (e.g., resulting in empirical fragility func- atures, higher loading rates, and high stress constraint inhibit plastic
tions). Lower-scale approaches utilize continuum stress or strain flow and increase stress, increasing propensity toward cleavage.
fields (rather than global deformations) and form the disciplinary Conversely, higher temperatures, slow loading rates, and low stress
core of fracture mechanics. Within this disciplinary core, at one end constraint promote plastic flow, favoring microvoid growth and co-
of the spectrum is traditional fracture mechanics where parameters alescence. The Charpy V-notch (CVN) temperature transition curve

© ASCE 03116001-2 J. Struct. Eng.

J. Struct. Eng., 03116001


Intact Microvoid nucleation and Coalescence
matrix growth and fracture
Dimpled fracture surface

0.1 mm
(a)
Crack propagation
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

Grain A

Grain B
Rupture through grains
River patterns and shiny facets
(b)

Fig. 1. Micromechanisms of fracture and corresponding failure surfaces: (a) microvoid growth and coalescence; (b) cleavage; transgranular cleavage
shown

Dynamic predate the advent of the discipline of fracture mechanics, often


Static Curve
Curve
considered coincident with the introduction of the Griffith (1921)
Absorbed energy

criterion for fracture, based on strain energy release rate. The ap-
Upper shelf – plication of these methods involves calculating a force or deforma-
microvoid growth
Temperature shift tion demand through structural analysis and then comparing it to a
corresponding capacity, determined through (ideally full-scale)
Transition region –
combination of microvoid component testing. This approach is accurate (assuming the tested
growth and cleavage components are sufficiently similar in scale and construction to the
Lower shelf – field prototypes), but expensive and difficult to generalize to dis-
predominantly cleavage similar structure components. Nevertheless, it is used when (1) there
is a low tolerance for inaccuracy or modeling uncertainty that may
Test temperature be introduced due to lower-scale methods, and (2) sufficient resour-
ces are available for large-scale testing. Successful applications of
Fig. 2. Characteristic temperature transition response of structural this approach include the multiuniversity campaign of large-scale
steels testing and analysis during the SAC steel project (Malley 1998),
motivated by the Northridge moment frame fractures. The design
considerations arising from these were based primarily on the ex-
perimental observations. Accordingly, standards such as AISC 358
well known to structural engineers is a direct macrostructural mani- (AISC 2010a) reference prequalified connections (FEMA 2000)
festation of this. Fig. 2 shows a typical CVN curve, plotting the that ensure similarity between the tested specimens and archetype
absorbed CVN energy (as a measure of toughness) against temper- connections, recognizing that purely experimental observations are
ature. Referring to the figure, three distinct regions are observed. At difficult to generalize. Although these methods are important, their
low temperatures, fracture is controlled by cleavage resulting in low intellectual basis is self-evident. This section focuses on the
toughness. At intermediate temperatures, a transition region is ob- specialized fracture mechanics approaches.
served, where some ductile tearing may precede cleavage, resulting
in increased toughness. Finally, at higher temperatures an upper
shelf is observed, where fracture is controlled entirely by ductile Traditional Fracture Mechanics
tearing. Fig. 2 also shows the effect of loading rate on CVN, such The fundamental concern of traditional fracture mechanics is
that an increased loading rate lowers toughness, especially in the precracked bodies. Reasons for this are practical as well as aca-
transition region. With this background, the next section discusses demic. From a practical standpoint, fracture is often triggered
fracture simulation methodologies for steel structures. around sharp cracks and discontinuities in structures (and less often
in smooth details). Academically, sharp cracks are interesting be-
cause they imply a stress singularity (stress and strain approaching
Models for Predicting Fracture in Steel Structures infinity) at the crack tip (Fig. 3), thereby precluding any stress- or
strain-based failure criterion. To overcome this problem, traditional
Empirical methods to characterize fracture (or more generically, fracture mechanics formulates the fracture criterion in terms of
structural failure) have been prevalent for several decades and parameters that characterize the stress field in the vicinity of the

© ASCE 03116001-3 J. Struct. Eng.

J. Struct. Eng., 03116001


K I ≥ K IC or J I ≥ J IC ð2Þ

In the preceding inequality, the left-hand side (K I or J I ) repre-


Increasing KI sents the toughness demand in a structural component, which is
σ
determined through finite-element simulations or sometimes (for
more regular geometries) through analytical or regressed formulas
[refer to Rooke and Cartwright (1976) for K I solutions and Kumar
et al. (1984) for J I solutions]. The right-hand side (K IC or J IC ) is
assumed to be a material property determined through standardized
r
testing [ASTM E399 (ASTM 2013a) for K IC, ASTM E1820
Singular stress fields (ASTM 2013c) for J IC ]. This departure from continuum-based
failure theories (which are based on pointwise stress evaluations)
is necessitated by the stress singularity. In reality, the stresses (and
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

strains) at the crack tip are regularized (meaning they are not equal
to infinity) due to (1) the formation of a plastic zone around the
crack tip, (2) microstructural features that invalidate continuum me-
chanics over these small scales, or (3) finite crack widths. Even
today, it is computationally expensive to simulate these features
Fig. 3. Crack tip singularity implicit in linear elastic fracture mechanics with the aim of characterizing stress and strain fields in the close
vicinity of the crack tip. The usefulness of a far-field or remote
parameter such as the stress intensity factor is even more under-
standable considering the computational resources available when
crack, rather than a discrete location within the body. The implicit LEFM and EPFM were first developed.
assumption is that these parameters are correlated with crack tip Traditional fracture mechanics is valid only as long as the tough-
stress and strain fields that activate the microstructural processes ness parameters retain a unique correspondence with the stress-
described previously. Two popular parameters associated with tra- strain fields in the vicinity of the crack tip, even if these fields
ditional fracture mechanics are the stress intensity factor (denoted are not consistent with the underlying theory (e.g., linear elasticity)
K I ), and the contour J-integral (denoted J I ). The K I forms the used to formulate the parameters. In general, this is true when the
core of linear-elastic fracture mechanics (LEFM), whereas the size of the plastic zone at the crack tip is significantly smaller than
latter is arguably the most important parameter within elastic- the region of K I or J I dominance (i.e., the region where the singular
plastic fracture mechanics (EPFM). Other EPFM parameters such stress-strain fields consistent with K I or J I dominate the overall
as the crack tip opening displacement (CTOD) or crack tip open- stress field). In practical terms, this results in size limitations on
ing angle (CTOA) are not discussed here, albeit extensive discus- calibration specimens for K IC and J IC and corresponding size lim-
sion on these may be found in textbooks (e.g., Anderson 1995; itations for structural components within which they may be used to
Saxena 1997). predict fracture. For example, LEFM is valid if the characteristic
Both K I and J I may be interpreted as fracture toughness de- dimension (e.g., flange thickness, crack length) of a component
mand parameters determined from the stress field in a loaded struc- is greater than approximately 2.5 × ðK IC =Fy Þ2. This indicates over-
tural component. The former (K I ) is interpreted as the amplitude all elastic response before fracture. EPFM is valid when the speci-
of the singularity of the stress field near the crack tip in an elastic men size is greater than 10 × ðJ IC =Fy Þ, which typically allows for
body, shown in Fig. 3. Eq. (1) indicates that (1) the stress field may ductile fracture at the macroscale. The size limitations for EPFM
be expressed as a series expansion in terms p offfiffiffithe distance r ahead are more liberal than those for LEFM because EPFM formulations
of the crack tip, such that K I scales the 1= r term, which repre- admit material nonlinearity, thereby preserving the correspondence
sents the singularity because σ → ∞ as r → 0; and (2) a higher K I between J I and crack tip stress fields for a higher level of plasticity.
(shown by the dashed line in Fig. 3) implies higher stresses remote Because linear elasticity is a special case of nonlinear elasticity,
from the crack trip, although both curves approach infinity near the EPFM is always valid when LEFM is, but not vice versa, meaning
crack tip J I -based fracture assessment may be conducted even when the
response is expected to be linear elastic or brittle (following the
KI preceding size limitations). On the other hand, K I -based assess-
σ ¼ pffiffiffiffiffiffiffi
ffi þ higher order terms ð1Þ ment may be used only in the case of elastic or brittle fracture.
2πr A corollary of this is that for linear elastic response, J I and K I are
equivalent. In fact, they may be uniquely related such that J I ¼
The J-integral (J I ) = central parameter of EPFM. This param- K 2I × ð1 − ν 2 Þ=E. An analogous relationship may be used to relate
eter [proposed by Rice (1968)] may be interpreted as the strain en- K IC and J IC . In summary, traditional fracture mechanics ap-
ergy release rate per unit area extension of a crack in a nonlinear proaches may be used when (1) there is a sharp crack or flaw
elastic body. Similar to K I , J I may be interpreted as a stress inten- in the structural component, (2) the spatial extent of yielding is
sity parameter as well, as demonstrated by Hutchinson (1968) and small with respect to component size, and (3) the loading is
Rice and Rosengren (1968), defining the amplitude of a crack tip monotonic.
singularity. Thus, traditional fracture mechanics (LEFM or EPFM) Following the previous discussion, traditional fracture mechan-
rely on the notion that the amplitude of the stress field singularity ics [Eq. (2)] require (1) either regressed formulas or finite-element
defined by K I or J I (rather than pointwise stress values themselves) simulation to determine K I or J I in the structural component, and
may be used as a fracture parameter, such that both toughness de- (2) laboratory tests to calibrate K IC and J IC . Commercial and aca-
mands and capacities may be expressed in terms of this parameter. demic finite-element programs (e.g., Abaqus; Dodds et al. 2012)
Following this, the fracture criteria in traditional fracture mechanics are able to simulate stress and strain fields required for the deter-
may be expressed as mination of K I or J I subject to two considerations. First, the

© ASCE 03116001-4 J. Struct. Eng.

J. Struct. Eng., 03116001


modeled with zero width. Fig. 4(b) shows an alternative, in which
the crack is modeled as a finite width gap, with refinement of the
mesh to capture stress and strain gradients. The latter approach is
acceptable if the crack width is significantly lower than the critical
CTOD (∼0.25 mm for structural steels). Modern software plat-
forms such as Abaqus and WARP3D provide on-board capabilities
for K I and J I computations.
Laboratory tests for the determination of K IC and J IC are
informed by the standards ASTM E399 (ASTM 2013a), and
Rosette of “Quarter point” crack tip elements ASTM E1820 (ASTM 2013c), respectively, requiring sharp-
(a) cracked specimens such as the compact tension specimen shown
in Fig. 5(a). These specimens must conform to strict geometrical
Crack

limits to ensure small-scale yielding (SSY). Additionally, these


specimens require the application of fatigue loading prior to testing
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

to produce a zero-width precrack. These processes add significant


effort and cost to fracture toughness characterization. Owing to
these costs, standard K IC and J IC tests are rarely conducted for civil
structures (neither for materials specification nor for performance
assessment), and most design standards reference the CVN fracture
toughness, which may be determined more economically through
impact testing. For example, the AISC seismic provisions [AISC
341-10 (AISC 2010b)] require minimum material toughness for
Highly refined Crack tip elements 0.005 mm
various scenarios in terms of CVN. While these requirements es-
(b) tablish a baseline level of material toughness, they cannot be used
for quantitative fracture prediction. This necessitates the use of cor-
Fig. 4. Meshes for sharp crack simulation: (a) zero width crack with
relations between K IC and CVN to substitute direct measurements
quarter-point elements; (b) finite width crack with highly refined crack
of K IC . One such correlation, proposed by Barsom and Rolfe
tip mesh
(1969) is shown in Eq. (3)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K ID ¼ 0.646 × CVN × E ð3Þ
material constitutive response (von Mises plasticity for steel) must
be calibrated accurately, especially for high strains (greater than In this equation K ID is the critical stress intensity factor at a
0.5, i.e., 50% strain). Second, the finite-element mesh in the vicin- dynamic rate of loading (strainrate ≈ 105 mm=mm=s). This is sig-
ity of the crack tip must simulate the high stress and strain gradients nificantly lower compared with K IC for seismic or static loading
in this region. Fig. 4 shows examples of two acceptable mesh pat- (strainrate ≈ 0 − 0.1 mm=mm=s), implying that using the correla-
terns. Fig. 4(a) shows a crack tip rosette in which the element for- tion is conservative. However, if the full Charpy transition curve
mulations are enriched by moving the center node to the quarter (similar to that previously shown in Fig. 2) is available, then the
point toward the crack tip; in this manner, the elements can reflect correlation of Eq. (3) may be used along with a temperature shift
the singular stress field near the crack tip. In this case, the crack is (Fig. 2) to generate an equivalent curve for K IC. It is also pertinent

“Atomically Sharp”
fatigue precrack

θN

45°

(a) (b) (c) (d)

Fig. 5. Specimens commonly used for fracture model calibration: (a) compact tension specimen for K IC, JIC ; (b) circumferentially notched tension
(CNT) specimen; (c) inclined notch butterfly (IN) specimen; (d) rectangular notched (RN) specimen

© ASCE 03116001-5 J. Struct. Eng.

J. Struct. Eng., 03116001


Yield ULCF Low T history effects Cleavage
to mention here the Master Curve approach based on the work by

Yesd
Yes
Yes

Yes
No

No
No

No

No
Wallin (1998), which establishes the temperature dependence of
fracture toughness in an efficient manner through a limited number

Ability to simulate fracture under


of K IC tests. The Master Curve approach is standardized as per

Loading
ASTM E1921 (ASTM 2013b).

Yes
Yes

Yes
No
No

No

No

No
No
Local Approach to Fracture
Traditional fracture mechanics requires the correspondence of

Yes

Yes
No
No

No

No
No

No
No
its parameters with local stress fields under specified conditions
(small-scale yielding, monotonic loading). It is ineffective when
these conditions are violated. In contrast, local approaches con-

Yes

Yes
No
No

No

No

No

No
No
struct fracture criteria directly in terms of the stress and strain fields
(at the crack tip or in the zone of high distress). By relying on the
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

Yesc
Yesc

Yes

Yes
Yes

Yes

Yes

No
No
local stresses, these approaches are less constrained by require-
ments of specimen or component size, yielding, or the necessity of

sharp cracked for length scale

CNT, inclined notch butterfly


a sharp crack or flaw, and therefore are more generally applicable.

Sharp-cracked CTa, SENBb

tension (CNT) along with


Circumferentially notched
On the other hand, they are contingent on effective simulation of

(IN), rectangular notched

Sharp cracked similar to


(RN) specimens (Fig. 5)
Specimens (Fig. 5)
the stress and strain fields over extremely small scales (on the order
of steel grain size ≈ 0.1 mm). In turn, this presents challenges in

LEFM or EPFM
terms of computational demand and simulation of material constit-
utive response. Additionally, the notion of material homogeneity is

calibration
and others
also somewhat questionable at this scale, sometimes necessitating
the simulation of heterogeneous microstructural features such as
individual grains.
Unlike traditional fracture mechanics, local approaches seek to

but best practices developed


simulate specific mechanisms of fracture. The assumption is that

No established standards;
J IC , CTODC , CTOAC E1820 (ASTM 2013c),
fracture will initiate and propagate in accordance with the mecha- Calibration

ASTM standards

E1290 (ASTM 1999),


E2472 (ASTM 2012)
nism (microvoid growth or cleavage) that is more favorable based
on stress state, loading rate, and temperature. Because these mech-
anisms are dependent on material microstructure, numerous local

in research
models have been developed for specific materials; a comprehen-
sive review is available in Besson et al. (2006). In this paper, popu-
E399

lar models for two common mechanisms (microvoid coalescence


and cleavage) are presented. Table 1 summarizes the various local
models discussed in this section, along with their features, param-

B1 , B2 , B3 , B4 , n
eters that require calibration, and experiments required to cali-

Only if plastic zone is small, i.e., specimen or component size limitations are satisfied.


SWDM , λ, l
VGM , λ, l
Parameters


A1 , A2 , l

brate them.
l

σ f , λ
σu , m
,
K IC

Dcritical
VGM
Dcritical

Dcritical
Local Fracture Criteria for Microvoid Growth and
Coalescence
Void growth and coalescence around sulfide or carbide particles
is a primary fracture mechanism in structural steel. Criteria used
Eqs. (7) and (8) Microvoid growth, collapse,

Microvoid growth, collapse,


Micromechanism simulated

to describe this response are typically based on understanding of


microvoid or cavity growth in elastoplastic continua. Early ana-
coalesecence, and shear
Microvoid growth and

lytical work by McClintock (1968), and Rice and Tracey (1969)


Not associated with

Microvoid growth,

coalescence, shear

shows derivations for the cases of cylindrical and spherical


micromechanism

cavities, respectively. Their findings are similar, and are expressed


coalescence

coalescence

in Eq. (4)
Cleavage
Table 1. Summary of Selected Fracture Criteria

dR=R ¼ C1 × expðC2 × TÞ · dεp ð4Þ


Including spatial sampling of weak links.
Eqs. (5) and (6)

Referring to Eq. (4), the main observations are that (1) incremen-
Equations

Eq. (11)

Eq. (12)

Eq. (10)

Eq. (13)
Eq. (15)
Eq. (2)
Eq. (2)

tal void growth dR=R is proportional to the incremental equivalent


plastic strain dεp , and (2) the proportionality is a strong function
Single-edge notched bending.

of the stress triaxiality T ¼ σm =σe, where σm and σe are the mean


(or hydrostatic), and effective (or von Mises) stress, respectively.
The original derivations by Rice and Tracey (1969) suggest a value
Bao and Wierzbicki
Cyclic void growth
Void growth model

Ritchie Knott Rice

Compact tension.

of 0.283 for C1 and 1.5 for C2. The validity of this relationship
Stress-weighted
Stress-modified

has been demonstrated through numerical void cell simulations


damage model
critical strain

(Faleskog et al. 1998; Cooke 2015, Fig. 6) and through numerous


Criterion

Beremin

studies that have explored the response of cavities (or microvoids)


LEFM
EPFM

model

and their effect on ductile rupture under a range of stress states,


loading regimes, and material microstructures. Some examples
b

d
a

© ASCE 03116001-6 J. Struct. Eng.

J. Struct. Eng., 03116001


Multivoid cells to
characterize intervoid
interactions and localization

Coupon scale simulations for Single void cell to develop


Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

stress state determination fundamental void mechanics

Fig. 6. Void cell simulations in support of micromechanics-based fracture criteria

include work by Srivastava and Needleman (2012) examining void The use of coupled models (which are essentially constitutive
growth and collapse in monocrystals and Basu and Benzerga models with softening material response) results in mesh sensitivity
(2015) investigating the loading path dependence of void growth of the solution, making them somewhat problematic for application
controlled ductile fracture. Other refinements include incorporation in civil or structural simulations, where resources for mitigation of
of cavity shape effects on fracture, e.g., Pardoen and Hutchinson mesh sensitivity [e.g., the use of regularization schemes (Enakoutsa
(2000). These represent only some of the advances in a field that et al. 2007)] may be inaccessible or unfeasible. Accordingly, the
continues to evolve rapidly. Cumulatively, these studies have re- more accessible uncoupled models are discussed in this paper. In
sulted in (and continue to refine) insights regarding dependencies this context, recent efforts by Kiran and Khandelwal (2014) to de-
of void response on key aspects of the stress-strain fields and their velop weakly coupled models that balance the accuracy of coupled
histories. Consequently, the large majority of local fracture models models with the convenience of uncoupled ones are particularly
are based on these insights. Due to the multiplicity of conditions for notable.
which these insights have been developed, variants of local models Within the damage mechanics category, the void growth model
that rely on them are similarly numerous. This paper focuses on a (VGM) is relatively well known. This model, which has been used
subset of these that have been recently examined and applied to by Panontin and Sheppard (1995) and Kanvinde and Deierlein
structural steel components. (2006), among others, as a fracture criterion predicts fracture when
Local fracture criteria that seek to simulate microvoid-induced the following condition [based on a direct integration of Eq. (4)] is
ductile fracture based on the previous research may be broadly clas- satisfied:
sified into two categories: (1) damage mechanics–based criteria, in Z
which damage (which may be interpreted as proxy for void size) is lnðR=R0 Þ ≥ C1 × expðC2 × TÞ · dεp ≥ lnðRcrit =R0 Þ ð5Þ
loading
evaluated as a history variable quantity based on evolving stresses
or strains, for example, by integrating the left-hand side of Eq. (4); Simplifying further, this may be expressed as
or (2) state-based criteria, in which a critical fracture strain is
expressed as a function of stress state. Within the damage mechan- DVGM ≥ Dcritical
VGM ð6Þ
ics category, the models may be further classified as coupled or
uncoupled: In the previous equation, the void growth damage DVGM ¼
• Coupled models are able to simulate interactions between ∫ loading expðC2 × TÞ · dεp , and its critical value Dcritical
VGM ¼
the fracturing material and the surrounding stress field by incor- lnðRcrit =R0 Þ=C1 . Simplification in this manner allows for the
porating damage as a quantity that affects material softening. calculation of the void growth damage as a demand parameter
A classic example is the Gurson (1977) model, which based solely on simulation [C2 is often considered to be a material-
simulates material softening response due to void growth. independent parameter with the value 1.5 (Kanvinde and Deierlein
Numerous coupled damage mechanics models have been 2006)], whereas Dcritical
VGM may be interpreted as a material property
subsequently developed to target specific aspects of response, (or a damage capacity) requiring calibration. For structural steels,
which are triggered under different stress states or material Dcritical
VGM is in the range of 1.00–5.00. Kanvinde and Deierlein (2007)
microstructure. For example, Kiran and Khandelwal (2015) extended the void growth model to ULCF, which typically occurs
incorporate the effect of void elongation along with dilation during seismic loading. Unlike high-cycle fatigue, which occurs in
in such a model, whereas others (e.g., Ruggieri et al. 1996) bridges or mechanical components (involving thousands or even
have incorporated material-specific length scales into such millions of cycles), ULCF is characterized by a small (less than 20)
models. number of cycles with high strain amplitudes (several times yield).
• Uncoupled models apply the fracture model as a failure criterion In terms of micromechanisms, ULCF bears a close resemblance
(through a postprocessing check) to stress and strain fields to fracture (because it involves cyclic growth and collapse of voids)
that are computed based on a constitutive model that does rather than high cycle fatigue [which involves the processes of
not simulate damage or softening. As a result, these are some- dislocation slip and decohesion (Suresh 1998)]. The cyclic void
what less generalizable as compared with the coupled models, growth model (CVGM) (Kanvinde and Deierlein 2007) has a dam-
especially if the strain distribution (or localization) leading to age functional form similar to the VGM, except it accounts for
fracture is driven by damage itself. cyclic deterioration, as in Eq. (7)

© ASCE 03116001-7 J. Struct. Eng.

J. Struct. Eng., 03116001


DCVGM ≥ Dcritical
CVGM ð7Þ center (where fracture initiates). This suggests that generalizing
these models to nonaxisymmetric stress states is potentially in-
In the preceding, the damage index DCVGM accounts for accurate; the SWDM addresses this. As summarized in Table 1,
reversed cyclic loading, such that the SWDM requires the calibration of two parameters (in addi-
Z Z tion to the length scale), similar to the CVGM. One is the
DCVGM ¼ expð1.5 × jTjÞ · dεp − expð1.5 × jTjÞ · dεp Dcritical
SWDM , and the other is λ (identical in interpretation and values
T≥0 T<0 as for the CVGM).
ð8Þ Other notable models in this category (uncoupled damage
mechanics models addressing low triaxiality) include those by
Eq. (8) separates the loading history into positive and negative Wen and Mahmoud (2016a) for monotonic loading and Wen and
triaxiality cycles, reflecting that voids grow and shrink during al- Mahmoud (2016b) for cyclic loading. These models are similar to
ternating cycles of positive and negative triaxiality. The damage the SWDM discussed previously, but propose somewhat different
capacity may be determined according to the following expression: functional forms for damage accumulation based on different data
sets. Similar developments have been made in the area of coupled
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

Dcritical critical
CVGM ¼ DVGM × expð−λ · εacc Þ ð9Þ damage mechanics models as well; for example, Nahshon and
Hutchinson (2008) present such a model for shear-dominated,
According to the preceding expression, the damage capacity de- low-triaxiality stress states.
creases with loading as a function of the accumulated plastic strain The models discussed previously require the numerical integra-
εacc . At the initiation of loading, εacc ¼ 0, and as a result Dcritical
CVGM ¼ tion of stress and strain quantities at each continuum location to
Dcritical
VGM with the implication that for monotonic loading the CVGM apply the fracture criterion. On the other hand, state-based models
is identical to the VGM. Consequently, as summarized in Table 1, are based on the instantaneous stress and strain fields (rather than
the CVGM requires the calibration of two parameters (in addition their histories). For example, a class of models [termed stress-
to a length scale parameter, discussed subsequently), one of which modified critical strain (SMCS)] take the following general form:
is the damage capacity of the monotonic VGM, i.e., Dcritical VGM ; the
other is λ, characterizing the rate of cyclic deterioration. For struc- εp ≥ εcritical
p ¼ A1 · expð−A2 · TÞ ð11Þ
tural steels, this parameter is in the range 0.0–1.0. Myers et al.
(2009) demonstrate an application of CVGM to large-scale struc- The preceding criterion (in which A1 and A2 are calibrated
tural tests. The original derivations by McClintock (1968) and Rice parameters) may be evaluated without numerical integration (as
and Tracey (1969) were developed for high triaxiality (T > 1.0) required in the VGM, CVGM, and SWDM models). In fact, the
owing to the importance of fracture at crack tips. Low triaxiality functional form shown in Eq. (11) arises directly through integra-
is commonly found at protruding corners of structures (e.g., tips tion of Eq. (5) for the VGM, assuming that the triaxiality is T, a
of beam flanges) or in high-shear conditions that are common constant over the loading process. Consequently, SMCS-type mod-
in civil structures. To address this, Smith et al. (2014) proposed els are accurate when the temporal variations in triaxiality are
the stress-weighted damage model (SWDM), whose fracture modest. The variants of the SMCS models are typically distin-
criterion is guished through the calibrated values of A1 and A2 ; the latter is
Z usually taken as 1.5. The parameter A1 is often the only material
DSWDM ¼ eλ·εacc × ðe1.3T − e−1.3T Þ · e0.5X · dεp ≥ Dcritical property that requires calibration. Popular variants of the SMCS-
SWDM
loading type models include those by Hancock and Mackenzie (1976),
ð10Þ Johnson and Cook (1983), and Rousselier (1987).
The SMCS models are state-based counterparts to the damage-
The preceding criterion is similar to the CVGM or the VGM in mechanics models of VGM or SWDM. Consequently, in addition
that it is based on the damage index DSWDM , which must be com- to their inaccuracy with respect to loading histories, they cannot
pared with its critical value Dcritical
SWDM . However, the SWDM criterion
accurately describe the effects of low triaxiality or the Lode param-
offers improvements compared with the CVGM in two respects: eter. Analogous state-based counterparts for low-triaxiality fracture
• It reflects a lower damage rate for low triaxiality such that at have been developed as well. Major contributions in this area have
zero triaxiality the damage rate [the integrand in Eq. (10)] is 0. been made by current and former members of the Wierzbicki
This is consistent with experimental data by Bridgman (1964) research group at the Massachusetts Institute of Technology. For
and Smith et al. (2014) and is not accounted for by the CVGM example, Bao and Wierzbicki (2004) presented a fracture criterion
and similar models. targeted toward crashworthiness assessment of vehicles, where
• It incorporates the effect of the deviatoric parameter (or Lode low-triaxiality stress states are frequently encountered. This cri-
angle parameter) X, which may be determined as X ¼ terion is similar to the SMCS in terms of its history independence
pffiffiffi
3 3 · J 3 =ð2 · J 3=2
2 Þ, where J 2 and J 3 are the second and third εp ≥ εcritical
p ¼ B1 · expð−B2 · TÞ − ½B1 · expð−B2 · TÞ
invariants of the deviatoric stress tensor. This parameter may be
interpreted as an additional indicator of stress state, such that − B3 · expð−B4 · Tފ · ð1 − X 1=n Þn ð12Þ
X ¼ 1 indicates an axisymmetric stress state (e.g., at the cen-
ter of a cylindrical bar) and X ¼ 0 indicates plane strain. This The parameters B1 , B2 , B3 , B4 , and n are calibrated material
parameter has been shown through experimental data (Bao and parameters. Fig. 7(a) shows the failure locus (i.e., the strain capac-
Wierzbicki 2004; Xue 2007) to strongly influence fracture. ity as a function of the stress triaxiality) for the SMCS model
Functional forms derived by Rice and Tracey (1969) and [Eq. (11), which does not consider the effect of the Lode param-
McClintock (1968) are developed only for the axisymmetric eter], as well as Eq. (12), which does. Referring to Fig. 7(b),
condition with X ¼ 1. Moreover, a majority of experimental Eq. (12) can capture the effect of the Lode parameter (meaning
specimens used to develop and calibrate models based on this it can distinguish between axisymmetric and plane strain situa-
functional form (e.g., the CVGM and VGM) are cylindrically tions even for the same stress triaxiality), whereas the SMCS-type
notched specimens with an axisymmetric stress state at the models cannot. However, similar to the SMCS models, these are

© ASCE 03116001-8 J. Struct. Eng.

J. Struct. Eng., 03116001


ε pfracture ε pfracture

Axisymmetric
stress state

Plane strain
Triaxiality T
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 7. Fracture locus for (a) SMCS-type models, invariant to deviatoric parameter [Eq. (11)]; and (b) Bao and Wierzbicki (2004) model accounting
for deviatoric parameter dependence [Eq. (12)]

 1=m
unsuitable if the stress state changes significantly over the loading
Z
history; in this case the damage mechanics–type models are more σw ¼ ð1=V 0 Þ σm
max · dV ð15Þ
Ω
appropriate. Some recent research that has focused on developing
and refining models in this general area includes Roth and Mohr
(2016), Bai and Wierzbicki (2010), and Xue (2007). In Eq. (15), V 0 = reference volume, which may be arbitrarily
chosen in concert with the material parameters σu and m. The
Local Criteria for Cleavage Fracture underpinnings of this approach and calibration methods are dis-
Cleavage fracture occurs due to the rupture of molecular bonds cussed at length in Beremin (1983), and by Matos and Dodds
within (or on the boundaries of) metal grains, with the implication (2001) in a structural engineering context (fracture of beam-
that it is stress (rather than strain) controlled. It is more probable column connections in the Northridge earthquake). In the context
when the material yield stress is increased either due to low tem- of this paper, it suffices to observe that (1) once V 0 , σu , and m
perature, high loading rates, or high stress constraint. Cleavage is have been selected, Eqs. (14) and (15) may be evaluated to es-
typically triggered around weak links (such as inclusions) in the timate the probability of cleavage fracture; and (2) this probability
material microstructure, and its onset is quite random, depending reflects the spatial sampling of presumed weakest links in the
on the manner in which the stress field samples these weak-links. material, interacting with a nonhomogenous stress field in the
Two popular local criteria are discussed here: one is commonly loaded component.
termed the ritchie knott rice (RKR) model (Ritchie et al. 1973), As discussed previously, cleavage fracture is preferred when
and the other is termed the Beremin model (Beremin 1983). the stresses are elevated due to a combination of constraint, tem-
The fracture criterion for the RKR model is a simple stress check, perature, and loading rate. For most structural components, as-
as shown in Eq. (13) suming that the loading rate and temperature are relatively
constant during loading, fracture often initiates due to the micro-
σmax ≥ σf ð13Þ void coalescence process (especially if sharp cracks are absent).
However, as loading progresses, stresses rise as a result of two
In this equation, σmax = maximum (tensile) principal stress at phenomena: (1) formation of crack surfaces due to microvoid
any location, and σf = fracture stress. As per the RKR model, the growth, resulting in an increase in stress triaxiality; and (2) strain
previous condition must be satisfied over a characteristic distance hardening. This increases the probability of triggering cleavage
(sometimes denoted λ ) to ensure adequate sampling of material fracture. Moreover, because cleavage fracture is not associated
weak links. The RKR model is deterministic in its formulation, with additional plastic dissipation, transition to cleavage usually
but acknowledges randomness in the cleavage process through the implies unstable propagation, as observed in many buildings after
sampling parameter. The Beremin model (Beremin 1983), popu- the Northridge earthquake.
larly known as the Weibull stress method, formalizes this random- The local models described previously are applied through
ness in a rigorous way, such that the fracture criterion is described stress and strain fields determined through continuum simulations.
in a probabilistic manner, wherein the probability of fracture may This means that in theory they can be applied at any continuum
be expressed as point, which has zero volume. This interpretation is meaning-
less unless there is a finite volume associated with this point
because fracture occurs only when microstructural features (such
Pf ¼ 1 − exp½−ðσw =σu Þm Š ð14Þ
as void inclusions, which have finite dimensions) are activated.
The Beremin model addresses this explicitly. For the void growth
In this equation, σu and m are material parameters (sometimes models, a characteristic length (or volume) parameter must often be
termed Weibull moduli), whereas the Weibull stress σw is a demand included in the model definition; this length is on the order of 2–10
quantity; more specifically, an averaged stress causing a consistent times the material grain size. Once this length scale is established,
probability of failure accounting for weakest-link sampling over the any of the model criteria described Table 1 may be implemented
entire component. It is determined as follows: based on stress and strain fields determined from appropriately

© ASCE 03116001-9 J. Struct. Eng.

J. Struct. Eng., 03116001


Increasing loading Application to Common Structural Engineering
parameter, such as Situations
deformation or J-integral
D / Dcrit This section examines the suitability of various models (discussed
in the previous section) for common structural engineering scenar-
ios. The discussion is premised on the notion that the ultimate ob-
Solid line indicates stress field jective of fracture modeling is to determine a global (or structural)
deformation corresponding to force or deformation at which fracture (and associated loss of
fracture strength) occurs. This may be required for a component that is
being designed or is already in service, or sometimes for develop-
ing broader insights into effects of detailing or material properties
in support of design standards. Referring to prior discussion, differ-
ent fracture models require disparate resources for simulation as
well as material characterization, resulting in varying levels of ac-
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

curacy for different applications. Following this, the primary objec-


tive of this discussion is not to advocate for a particular modeling
approach, but rather to provide perspective for optimal modeling
l r = Distance ahead of
strategies, recognizing that the optimal strategy is influenced by
crack or notch tip
available resources for calibration experiments and computational
Fig. 8. Determination of material length scale or volume parameter l resources, as well as the desired accuracy in prediction, noting that
a full-scale test of the component is the most accurate but also likely
the most expensive approach. The section describes conditions
commonly occurring in structural engineering and examines vari-
constructed finite-element simulations and calibrated material ous modeling options for each. Fig. 9 schematically illustrates
parameters. Fig. 8 shows this approach schematically. structural components where these conditions may be encountered.
Because the local models are an active area of research Referring to Fig. 9, the primary classification of the scenarios is
(with a multiplicity of functional forms and so on), methods whether they include sharp cracks or smooth transitions in geom-
to calibrate their parameters are, in general, not standardized. etry. Subclassifications pertain to whether the components are mon-
Consequently, calibration is based on (1) experimental coupons otonic or cyclic, and whether there is the expectation of large-scale
(such as cylindrically notched tension coupons with varying yielding prior to fracture. These are not all possible combinations,
notch dimensions, Fig. 5) to interrogate various stress states, and primarily because some (e.g., blunt notched conditions with large-
(2) complementary finite-element simulations. The stress-strain or small-scale yielding) are essentially equivalent in terms of their
data from the finite element model corresponding to the instant amenability to modeling approaches. Before discussing each sce-
of fracture observed in the experiments is then used to back- nario, it is useful to establish the connotations of sharp crack versus
calibrate the parameters of the local model by enforcing the frac- smooth transitions and small- versus large-scale yielding.
ture criterion at this instant. Some studies (e.g., Myers et al. Components may be classified as sharp cracked if they contain
2010, 2014) have focused exclusively on methods for parameter cracks (or flaws) that are introduced without removal of material
estimation). Table 1 and Fig. 5 summarize calibration specimens (sometimes referred to as atomically sharp cracks). Such cracks are
for various models. typically the result of high cycle fatigue where cracks grow under

High (sharp cracked) Intermediate (notches, holes local Low (corners and shear bands)
buckles )
Stress triaxiality
Monotonic

Cyclicity

Partial Joint Penetration with unfused Shear bands in block shear


Net section fracture around holes
weld root
Seismic
ULCF

Seismic beam column connection with


Cyclically buckling brace with local Seismic column base plate connection
flaw at weld root
High Cycle Fatigue buckling induced fracture with corner fracture
(Vehicular/ mechanical)

Fig. 9. Common structural engineering situations in the space of stress triaxiality and load cyclicity

© ASCE 03116001-10 J. Struct. Eng.

J. Struct. Eng., 03116001


relatively low-amplitude cyclic stresses through the microstructural Scenario 1b: Sharp-Cracked Component Subjected to Cyclic
processes of slip and decohesion of dislocations (Suresh 1998). Loading
Cracks in steel bridge decks are a common civil engineering exam- In this case, the geometric configuration is similar to Scenario 1a,
ple. Sharp cracks may also be formed when fracture initiates at except that the loading is cyclic. In this situation, the nature of
a blunt notch or in a smooth structural detail. Some types of flaws cyclic loading is the primary factor in selecting the optimal mod-
(e.g., those produced by lack of fusion in welded components) or eling strategy. If the situation can be categorized as high cycle
in partial joint penetration (PJP) welds may be considered sharp for fatigue (a bridge deck under vehicular loading or a mechanical
practical purposes, although they are not atomically sharp in theory. component), then fatigue crack growth may be calculated through
Following work by McMeeking and Parks (1979), if the crack established procedures such as the Paris fatigue law (Paris and
width is significantly less than the CTOD, which is on the order of Erdogan 1963). For every computed crack length and applied load
0.25 mm for low-carbon structural steels, of a material, then it may (either an overload or the peak tensile load of the cyclic loading),
be conservatively treated as sharp for the purposes of fracture fracture may be evaluated by assuming this to be equivalent to
prediction. Scenario 1a (i.e., a precracked body under monotonic loading),
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

The notion of SSY is interpreted in the context of its effect on with the final cycle being considered as the monotonic load. With
the accuracy of traditional fracture mechanics; it is generally under- the context of Scenario 1a, traditional EPFM and LEFM or local
stood to mean that the plastic zone is well confined (typically in the approaches may be applied (as discussed in the previous section),
vicinity of the crack tip) such that traditional fracture mechanics depending on the level of anticipated yielding.
parameters (K I , J I , or CTOD) bear unique correspondence with If the component is subjected to earthquake-type ultralow
crack tip stresses. Thus, SSY for structural components may be cycle fatigue, then large-scale yielding is anticipated even during
defined because it is for the validity of standard K IC or J IC tests. the cyclic loading, with a small (<20) number of cycles to failure.
With this background, various structural engineering scenarios are This situation may be interpreted as one of fracture due to cycli-
discussed. cally induced damage, rather than one of fatigue crack growth.
In this case, local criteria such as SWDM [Eq. (10)] or CVGM
[Eq. (8)] may be applied in a manner similar to the VGM for
Scenario Type 1: Structural Components with Scenario 1a.
a Sharp Crack
For structural components with a sharp crack, two subscenarios are
Scenario Type 2: Structural Components without a
distinguished by whether the component is loaded monotonically
Sharp Crack
or cyclically. Fig. 9 shows a PJP welded connection with a lack-
of-fusion flaw, and a pre-Northridge moment connection with a The possibility of fracture without the presence of a sharp crack is
pre-existing flaw, both of which may be considered structural com- of special interest in civil structures. This may occur in zones of
ponents with a sharp crack. stress or strain concentration around holes or blunt notches, in
zones of local buckling, or simply in smooth structural members
Scenario 1a: Sharp-Cracked Structural Component subjected to large inelastic strains. The selection of modeling
Subjected to Monotonic Loading strategy for structural components without a sharp crack is more
An example of this is a PJP welded butt splice (Fig. 9). In these straightforward than for cracked components because traditional
types of situations, the appropriate fracture model is motivated by fracture mechanics are invalidated at the outset due to the absence
the extent of yielding. The specimen size limitations for the use of a sharp crack and the high likelihood of large-scale yielding.
of linear-elastic fracture mechanics [as outlined in ASTM E399 Moreover, without the presence of a sharp crack (and the asso-
(ASTM 2013a)] or elastic-plastic fracture mechanics [as outlined ciated increase in stress triaxiality) it is unlikely that fracture will
in ASTM E1820 (ASTM 2013c)] may be used as general guide- initiate due to cleavage (recall that high triaxiality raises the prin-
lines to assess the size of the structural component (given material cipal stresses that are responsible for cleavage). This implies that
yield stress and fracture toughness) in the context of K IC or J IC local criteria for ductile fracture are usually applicable. Within
validity. If determined to be valid, standard LEFM or EPFM these local criteria, the specific criterion may be selected based
approaches (outlined in prior discussion) are suitable, where the on combination of factors including (1) nature of loading (mon-
toughness demand is determined from finite-element simulations otonic or ULCF), (2) stress state, defined by triaxiality and Lode
and the capacity is determined either through standardized testing parameter, and (3) evolution of stress-state through loading his-
or through correlations such as the one outlined in Eq. (3). tory, affecting whether history-based or state-based criteria are
In sharp-cracked structural components where traditional used. Criteria such as the SWDM are most general, being valid
fracture mechanics approaches are inapplicable (this is common for cyclic and monotonic loading, a range of stress states, and
in structural components with sharp but shallow cracks and high also loading history effects. However, these models also require
material toughness), local criteria are necessary. Specifically, this the calibration of additional parameters, which are redundant in
may include the history-based VGM [Eq. (6)] or its state-based specific scenarios (high triaxiality, monotonic loading). Notable
counterpart [Eq. (11)]. Chi et al. (2006) and Kanvinde et al. (2008) applications of these models include safety assessments of liquid
indicated that for these situations, either of these approaches is storage tanks under seismic loading (Prinz and Nussbaumer
more effective compared with traditional approaches. However, 2012) and bolted steel connections sensitive to block shear (Wen
because sharp-cracked situations involve steep strain gradients near and Mahmoud 2015).
the crack tip, the fracture predictions are sensitive to the character- Even in smooth components, the initiation of fracture (which
istic distance l , which must be calibrated accurately; the aforemen- may be predicted by a local criterion for ductile fracture) reflects
tioned references describe procedures to accomplish this. Because material separation and the formation of a sharp crack. Subsequent
cleavage fracture is always possible, it is also prudent to apply the to this, fracture propagates as a sharp crack, which typically in-
corresponding criterion (e.g., the Weibull approach or the RKR cludes some extent of ductile tearing followed by transition to
model) concurrently with the ductile fracture models. cleavage.

© ASCE 03116001-11 J. Struct. Eng.

J. Struct. Eng., 03116001


Limitations of the Current Framework and Areas for most well known of these. However, significant work has
Future Research been done in developing and refining these general ap-
proaches; these have been discussed previously in this paper
The limitations of individual fracture approaches (LEFM, EPFM, in the context of coupled damage mechanics models.
and local criteria) have been detailed in the foregoing sections. This • The use of cohesive zone models in which closing tractions
section focuses on issues within the modeling framework in its are applied on the crack faces behind the opened crack tip;
entirety. These broader issues impede the widespread application these provide regularization of the stress field, but suffer
of these well-developed models within the structural engineering from other limitations in terms of establishing fracture cri-
profession: teria. Elices et al. (2002) provide a comprehensive review of
1. Nonuniqueness of constitutive parameter fitting: Several models cohesive zone models.
outlined in this paper (especially the local models, which rely on 3. Heterogeneity and spatial randomness of material properties:
the local continuum strains) require finite-element simulations In structural components, the areas of high stress and strain (and
that must be able to simulate large strains (on the order of high stress-strain gradients), which occur at discontinuities such
0.5–1.0, i.e., 50–100%), which occur in highly localized regions as connections, often intersect areas of highly heterogeneous
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

of necking or local buckling. In turn, these finite-element simu- material properties due to welding. Welding introduces sharp
lations rely on multiaxial constitutive models (typically with gradients in material toughness, such that the weld and base
a von Mises yield surface and cyclic hardening properties). metal (with higher toughness) sandwich a low-toughness heat-
Calibration of these models for large strains (greater than 0.3 affected zone. This interaction creates additional challenges for
or so) cannot be performed using stress-strain curves from stan- simulation of fracture, especially using local criteria. A metho-
dard tests because the stress state in standard coupons becomes dology to assess fracture (with high confidence and tractability
nonhomogenous at lower strains (0.1–0.2) due to necking or of implementation) in these situations remains elusive.
buckling. Necessitated by this, calibration is often performed 4. Trade practices for specifying fracture toughness: The structural
by matching load-displacement curves of calibration specimens engineering profession relies primarily on the Charpy V-notch
to those obtained through complementary finite-element simu- energy measured at specific temperatures as the primary indica-
lations. In these situations, multiple parameter sets produce tor of steel toughness. From the perspective of fracture simula-
strain fields that match the measured load-displacement re- tion, this raises two issues: (1) the standards (e.g., AISC 2010b)
sponse, resulting in nonunique parameter fits. This means that establish only minimum requirements for fracture toughness,
multiple parameter fits may match the load-displacement curves such that the actual CVN energy is not known, and (2) even
accurately, albeit with different internal strain fields. Conse- assuming that the CVN is known, it cannot be conveniently con-
quently, the calibrated parameter set may not (in general) reflect verted to an equivalent K IC or J IC, for example, due to the un-
true material response, leading to large errors in determination certainty associated with the temperature shift (if the full CVN
of stress-strain fields, and consequently fracture. Cooke and curve is unavailable). The resulting uncertainty in the fracture
Kanvinde (2015) describe these issues. toughness results in conservative design, which while reassuring
2. Simulation of crack propagation: The primary focus on this from the standpoint of safety, is not compatible with accurate
paper is on fracture initiation, with the assumption that in most simulation of expected structural performance. In this regard,
structural engineering situations the initiation of fracture must it is also relevant to note that nominally identical steel samples
be prevented for acceptable performance. However, initiation [e.g., ASTM A572 (ASTM 2015)] often have large variation in
does not always result in propagation and failure, and may be fracture properties necessitating the use of material-specific
arrested depending on the material toughness, stress gradient toughness data for fracture assessment.
(crack may run into zone of compression), strain energy stored 5. Standardization of local approaches: The calibration methods
in the component, rate of loading, and the onset of cleavage for the local models and approaches to implement these models
fracture (which is quite random). Simulating crack propagation into research and advanced engineering practice have not been
requires the integration of fracture criteria along with numerical standardized. Nevertheless, best practices have been developed
schemes that can simulate this separation within finite-element for these, albeit within individual research projects; these best
models. This is challenging for many reasons, especially practices are described in preceding sections. In this regard, it is
because simulated crack propagation may be highly mesh- noted that guidelines for seismic performance assessment of
dependent in terms of crack path as well as the extent of pro- frame structures [e.g., FEMA P695 (FEMA 2009)] have greatly
pagation itself. Some common approaches for simulating crack accelerated the adoption of advanced simulation techniques into
propagation include the following: mainstream research and practice. Standardization of methods
• Introduction of free surfaces in the simulation model. This for fracture simulation in civil structures has the potential to
may be done explicitly, for example, through the insertion be similarly transformative.
of traction-free surfaces into the model topology, with atten- Even if fracture occurs during structural failure, fracture may
dant mathematical regularization, i.e., mitigation of stress not be the controlling event. For example, Fell et al. (2009) dem-
singularities that arise at a sharp crack tip (Baldwin and onstrated (for cyclically buckling steel braces) that local strains
Rashid 2013). Alternatively, this may be accomplished increase very rapidly with respect to applied global displacement
through the use of a field order parameter, which establishes after the onset of local buckling, with the implication that after this
a smooth transition between ruptured intact material phases instant, global displacement corresponding to fracture is virtually
and modifies material governing equations in a manner that insensitive to characterizations of local fracture strain (for example,
reflects crack initiation and propagation. Notable references through a local criterion). In such cases, it is more beneficial to
in this area include Miehe et al. (2016), and Raina and Miehe predict local buckling accurately rather than the fracture that usu-
(2016). Ambatti et al. (2015) provides a general summary of ally follows it. Similar situations may be encountered in tensile
phase field modeling, as this approach is known. loading, where the onset of necking amplifies strain, and predicting
• The use of a softening constitutive model, along with finite- this event (rather than the fracture that follows it) is more important.
element removal. The Gurson (1977) model is possibly the This suggests why, for several situations (such as net section

© ASCE 03116001-12 J. Struct. Eng.

J. Struct. Eng., 03116001


fracture of bolted connections), the use of the ultimate strength at Davis. The author is also grateful to sponsors of his research over
(associated with necking) is an excellent predictor of fracture. the years, including the National Science Foundation (Grants
In closing, it is informative to discuss the adoption of fracture #CMMI 0825155, 118634, and 143400) and the American Institute
mechanics into structural engineering practice. Although this began of Steel Construction. Recently, the author worked with Peter
with the SAC Steel Project nearly two decades ago, recent years Maranian of Brandow Johnston and Associates and Leonard
have seen an increasing acceptance of fracture mechanics into sim- Joseph of Thornton Tomasetti Engineers for the fracture mechanics
ulation frameworks and design provisions, as well as specific con- design of the Wilshire Grand Tower; he is grateful to these indivi-
struction projects. For example, recent guidelines for structural duals for stimulating discussions and collaboration. The opinions
performance evaluation [e.g., ATC 114 (ATC 2016), under prepa- expressed in this paper are solely of the author.
ration] include explicit fracture mechanics–based equations for
simulation of structural components such as beam-column connec-
tions or splices. Similarly, design standards such as the upcoming References
seismic provisions (AISC 2010b) allow the use of PJP welded spli-
ABAQUS [Computer software]. Dassault Systèmes, Waltham, MA.
ces, in part based on fracture mechanics simulations of Stillmaker
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

AISC. (2010a). “Prequalified connections for special and intermediate steel


et al. (2016). Fracture mechanics simulation was also heavily used moment frames for seismic applications.” AISC 358, Chicago.
for the design and material specification of key components as well AISC. (2010b). “Seismic provisions for structural steel buildings.” AISC
as the spire of the (currently under construction) Wilshire Grand 341, Chicago.
Tower in Los Angeles, which will become the tallest building Almar-Naess, A., Haagensen, P. J., Lian, B., Moan, T., and Simonsen, T.
on the West Coast of the United States upon completion in 2017. (1984). “Investigation of the Alexander L. Kielland failure—
Metallurgical and fracture analysis.” J. Energy Resour. Technol.,
106(1), 24–31.
Summary Ambatti, M., Gerasimov, T., and De Lorenzi, L. (2015). “Phase-field
modeling of ductile fracture.” Comput. Mech., 55(5), 1017–1040.
This paper provides an overview of the state of the art in fracture American Welding Society. (2015). “Structural welding code.” AWS
prediction for steel structures in civil infrastructure. This consists D1.1:2015, Miami.
Anderson, T. L. (1995). Fracture mechanics, 2nd Ed., CRC Press,
of models developed over decades for varied application scenarios
Boca Raton, FL.
and within different technological and computational settings, re- ASTM. (1999). “Standard test method for crack-tip opening displacement
sulting in the absence of a unifying methodology for fracture pre- (CTOD) fracture toughness measurement.” ASTM E1290, West
diction across all scenarios. In view of this, the mission of the paper Conshohocken, PA.
is to contextualize what often seems like a patchwork of discordant ASTM. (2012). “Standard test method for determination of resistance to
methodologies, with the ultimate objective of providing guidance stable crack extension under low-constraint conditions.” ASTM
for effective simulation of fracture in the structural engineering pro- E2472 12e1, West Conshohocken, PA.
fession. For this purpose, it is critical to develop an understanding ASTM. (2013a). “Standard test method for determination of reference
of the fracture micromechanisms themselves and the relationship of temperature, T o , for ferritic steels in the transition range.” ASTM
various models to these mechanisms. The various models include E399, West Conshohocken, PA.
ASTM. (2013b). “Standard test method for linear-elastic plane strain frac-
the traditional approaches of linear-elastic and elastic-plastic frac-
ture toughness K IC of metallic materials.” ASTM E1921, West Consho-
ture mechanics, which are phenomenological, and the more recent hocken, PA.
local approaches, which seek to simulate the micromechanisms at a ASTM. (2013c). “Standard test method for measurement of fracture
lower scale. The local approaches do not suffer from some limita- toughness.” ASTM E1820, West Conshohocken, PA.
tions of the traditional approaches, but are challenging to effec- ASTM. (2015). “Standard specification for high-strength low-alloy
tively implement within a practical setting. Optimal strategies for columbium-vanadium structural steel.” ASTM A572/A572M, West
predicting fracture in common structural scenarios are outlined, Conshohocken, PA.
recognizing trade-offs between the complexity and cost of im- ATC (Applied Technology Council). (2016). “Development of accurate
plementation and the accuracy of prediction. While major ad- models and efficient simulation capabilities for collapse analysis to
vances have been made in terms of fracture criteria for a range of support implementation of performance based seismic engineering.”
ATC 114, Redwood City, CA.
conditions (stress states, loading types, micromechanisms), the
Bai, Y., and Wierzbicki, T. (2010). “Application of extended Mohr-
framework itself has significant limitations. With advances in com- Coulomb criterion to ductile fracture.” Int. J. Fract., 161(1), 1–20.
putational technology, fracture mechanics continues to evolve to Baldwin, A. T., and Rashid, M. M. (2013). “A geometric model of deco-
address these limitations and other issues faced by the structural hesion in solid continua.” Int. J. Fract., 180(2), 205–221.
engineering profession. Complementarily, because the profession Bao, Y., and Wierzbicki, T. (2004). “On fracture locus in the equivalent
emphasizes accurate performance assessment rather than prescrip- strain and stress triaxiality space.” Int. J. Mech. Sci., 46(1), 81–98.
tive code-based design, it is anticipated that fracture mechanics will Barsom, J. M., and Rolfe, S. T. (1969). “Correlations between KIC and
meaningfully integrate into the design process. charpy V-notch test results in the transition-temperature range.” Impact
Testing of Metals, ASTM STP 466, West Conshohocken, PA, 281–302.
Barsom, J. M., and Rolfe, S. T. (1999). Fracture and fatigue control in
Acknowledgments structures: Applications of fracture mechanics, 3rd Ed., ASTM, West
Conshohocken, PA.
The author would like to thank previous and current students and Basu, S., and Benzerga, A. (2015). “On the path dependence of the fracture
collaborators whose work has contributed to this paper through locus in ductile materials: Experiments.” Int. J. Solids Struct., 71(2015),
79–90.
research or engaging discussions. These include Gregory Deierlein
Beremin, F. M. (1983). “A local criterion for cleavage fracture of a nuclear
of Stanford University, Benjamin Fell of California State Uni- pressure vessel steel.” Metall. Trans., 14A(11), 2277–2287.
versity, Sacramento, Andrew Myers of Northeastern University, Besson, J., Moinereau, D., and Steglich, D., eds. (2006). “Local approach
Christopher Smith of the National Institute of Standards and Tech- to fracture.” Euromech-Mecamat 2006, 9th European Mechanics of
nology, Ryan Cooke of Forell Elsesser Engineers, and Kimberly Materials Conf., European Mechanics Society, Moret Sur Loing,
Stillmaker and Vincente Pericoli of the University of California France.

© ASCE 03116001-13 J. Struct. Eng.

J. Struct. Eng., 03116001


Bridgman, P. W. (1964). Studies in large plastic flow and fracture, Harvard Malley, J. O. (1998). “SAC steel project: Summary of Phase 1 testing
University Press, Cambridge, MA. investigation results.” Eng. Struct., 20(4–6), 300–309.
Chi, W. M., Kanvinde, A. M., and Deierlein, G. (2006). “Prediction of duc- Matos, C. G., and Dodds, R. H. (2001). “Probabilistic modelling of weld
tile fracture in welded connections using the SMCS criterion.” J. Struct. fracture in steel frame connections. Part I: Quasi-static loading.” Eng.
Eng., 10.1061/(ASCE)0733-9445(2006)132:2(171), 171–181. Struct., 23(8), 1011–1030.
Cooke, R. J. (2015). “Micromechanical simulation of ductile fracture McClintock, F. A. (1968). “A criterion for ductile fracture by the growth of
processes in structural steel.” Ph.D. dissertation, Dept. of Civil and holes.” J. Appl. Mech., 35(2), 363–371.
Environmental Engineering, Univ. of California, Davis, CA. McMeeking, R. M., and Parks, D. M. (1979). “On criteria for J-dominance
Cooke, R. J., and Kanvinde, A. M. (2015). “Constitutive parameter cali- of crack-tip fields in large scale yielding.” Elastic Plastic Fracture,
bration for structural steel: Non-uniqueness and loss of accuracy.” ASTM STP 668, West Conshohocken, PA, 175–194.
J. Constr. Steel Res., 114, 394–404. Miehe, C., Teichtmeister, S., and Aldakheel, F. (2016). “Phase-field mod-
Dodds, R., et al. (2012). “WARP3D-release 17.6.0: 3-D nonlinear fracture eling of ductile fracture: A variational gradient-extended plasticity-
analysis of solids using parallel computers.” Structural Research Series, damage theory and its micromorphic regularization.” Philos. Trans.
No. 607, Univ. of Illinois at Urbana–Champaign, Urbana, IL. R. Soc. A., 374(2066), 20150170.
Elices, M., Guinea, G. V., Gomez, J., and Planas, J. (2002). “The cohesive Myers, A. T., Deierlein, G. G., and Kanvinde, A. M. (2009). “Testing and
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

zone model: Advantages, limitations, and challenges.” Eng. Fract. probabilistic simulation of ductile fracture initiation in structural steel
Mech., 69(2), 137–163. components and weldments.” Technical Rep. 170, John A. Blume
Enakoutsa, K., Leblond, J. B., and Perrin, G. (2007). “Numerical imple- Earthquake Engineering Center, Stanford Univ., Stanford, CA.
mentation and assessment of a phenomenological nonlocal model of Myers, A. T., Kanvinde, A. M., and Deierlein, G. G. (2010). “Calibration of
ductile rupture.” Comput. Methods Appl. Mech. Eng., 196(13–16), the SMCS criterion for ductile fracture in steels: Specimen size depend-
1946–1957. ence and parameter assessment.” J. Eng. Mech., 10.1061/(ASCE)EM
Faleskog, J., Gao, X., and Shih, C. (1998). “Cell model for nonlinear frac- .1943-7889.0000178, 1401–1410.
ture analysis—I. Micromechanics calibration.” Int. J. Fract., 89(4), Myers, A. T., Kanvinde, A. M., Deierlein, G. G., and Baker, J. W. (2014).
355–373. “A probabilistic formulation of the cyclic void growth model to predict
Fell, B. V., Kanvinde, A. M., Deierlein, G. G., and Myers, A. T. (2009). ultra low cycle fatigue in structural steel.” J. Eng. Mech., 10.1061/
(ASCE)EM.1943-7889.0000728, 04014028.
“Experimental investigation of inelastic cyclic buckling and fracture of
steel braces.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2009)135: Nahshon, K., and Hutchinson, J. (2008). “Modification of the Gurson
model for shear failure.” Eur. J. Mech., 27(1), 1–17.
1(19), 19–32.
Panontin, T. L., and Sheppard, S. D. (1995). “The relationship between
FEMA (Federal Emergency Management Agency). (2000). “Recom-
constraint and ductile fracture initiation as defined by micromechanical
mended design criteria for new steel moment-frame buildings.”
analyses.” Fracture Mechanics, Vol. 26, W. Reuter, J. Underwood, and
FEMA-350, Washington, DC.
J. Newman, eds., ASTM STP16379S, West Conshohocken, PA, 54–85.
FEMA (Federal Emergency Management Agency). (2009). “Quantification
Pardoen, T., and Hutchinson, J. W. (2000). “An extended model for void
of building seismic performance factors.” FEMA-P695, Washington,
growth and coalescence.” J. Mech. Phys. Solids, 48(12), 2467–2512.
DC.
Paris, P., and Erdogan, F. (1963). “A critical analysis of crack propagation
Griffith, A. A. (1921). “The phenomena of rupture and flow in solids.”
laws.” J. Basic Eng., 85(4), 528–534.
Philos. Trans. R. Soc. London, A, 221(582–593), 163–198.
Prinz, G. S., and Nussbaumer, A. (2012). “On the low-cycle fatigue capac-
Gurson, A. L. (1977). “Continuum theory of ductile rupture by void ity of unanchored steel liquid storage tank shell-to-base connections.”
nucleation and growth: Part I—Yield criteria and flow rules for porous Bull. Earthquake Eng., 10(6), 1943–1958.
ductile media.” J. Eng. Mater. Technol., 99(1), 2–15. Raina, A., and Miehe, C. (2016). “A phase field model for fracture in bio-
Hancock, J. W., and Mackenzie, A. C. (1976). “On the mechanics of ductile logical tissues.” Biomech. Model. Mechanobiol., 15(3), 479–496.
failure in high-strength steel subjected to multi-axial stress-states.” Rice, J. R. (1968). “A path independent integral and the approximate analy-
J. Mech. Phys. Solids, 24(2-3), 147–160. sis of strain concentration for notches and cracks.” J. Appl. Mech.,
Hutchinson, J. W. (1968). “Singular behavior at the end of a tensile crack in 35(2), 379–386.
a hardening material.” J. Mech. Phys. Solids, 16(1), 13–31. Rice, J. R., and Rosengren, G. F. (1968). “Plane strain deformation near
Johnson, G. R., and Cook, W. H. (1983). “A constitutive model and data for crack tip in power-law hardening material.” J. Mech. Phys. Solids,
metals subjected to large strain, high strain rate and high temperature.” 16(1), 1–12.
7th Int. Symp. on Ballistics, American Defense Preparedness Rice, J. R., and Tracey, D. M, (1969). “On the ductile enlargement
Association, Arlington, VA, 541–547. of voids in triaxial stress fields.” J. Mech. Phys. Solids, 17(3),
Kanvinde, A. M., and Deierlein, G. G. (2006). “The void growth model and 201–217.
the stress modified critical strain model to predict ductile fracture in Ritchie, R. O., Knott, J. F., and Rice, J. R. (1973). “On the relationship
structural steels.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2006) between critical tensile stress and fracture toughness in mild steel.”
132:12(1907), 1907–1918. J. Mech. Phys. Solids, 21(6), 395–410.
Kanvinde, A. M., and Deierlein, G. G. (2007). “A cyclic void growth model Rooke, D. P., and Cartwright, D. J. (1976). “Compendium of stress inten-
to assess ductile fracture in structural steels due to ultra low cycle sity factors.” Her Majesty’s Stationery Office (HMSO), London.
fatigue.” J. Eng. Mech., 10.1061/(ASCE)0733-9399(2007)133:6(701), Roth, C. C., and Mohr, D. (2016). “Ductile fracture experiments with
701–712. locally proportional loading histories.” Int. J. Plast., 79, 328–354.
Kanvinde, A. M., Fell, B. V., Gomez, I. R., and Roberts, M. (2008). Rousselier, G. (1987). “Ductile fracture models and their potential in local
“Predicting fracture in structural fillet welds using traditional and approach of fracture.” Nucl. Eng. Des., 105(1), 113–120.
micromechanics-based models.” Eng. Struct., 30(11), 3325–3335. Ruggieri, C., Panontin, T. L., and Dodds, R. H., Jr. (1996). “Numerical
Kiran, R., and Khandelwal, K. (2014). “Fast-to-computer weakly coupled modeling of ductile crack growth in 3-D using computational cell
ductile fracture model for structural steels.” J. Struct. Eng., 10.1061/ elements.” Int. J. Fract., 79(4), 309–340.
(ASCE)ST.1943-541X.0001025, 04014018. Saxena, A. (1997). Nonlinear fracture mechanics for engineers, CRC
Kiran, R., and Khandelwal, K. (2015). “A coupled microvoid elongation Press, Boca Raton, FL.
and dilation based ductile fracture model for structural steels.” Eng. Sheldon, L. (2010). “Structural failures: Testing lessons learned.” Test Eng.
Fract. Mech., 145, 15–42. Manage., 71(6), 10–12.
Kumar, V., German, M. D., Wilkening, W. W., Andrews, W. R., deLorenzi, Smith, C. M., Deierlein, G. G., and Kanvinde, A. M. (2014). “A stress-
H. G., and Mowbray, D. F. (1984). “Advances in elastic-plastic fracture weighted damage model for ductile fracture initiation in structural steel
analysis.” EPRI Final Rep. RP 1237-1, NP-3607, Electric Power under cyclic loading and generalized stress states.” TR 187, Blume
Research Institute, Palo Alto, CA. Earthquake Engineering Center, Stanford Univ., Stanford, CA.

© ASCE 03116001-14 J. Struct. Eng.

J. Struct. Eng., 03116001


Srivastava, A., and Needleman, A. (2012). “Porosity evolution in a Wen, H., and Mahmoud, H. (2015). “Prediction of block shear fracture
creeping single crystal.” Model. Simul. Mater. Sci. Eng., 20(3), in bolted connections.” 8th Int. Conf. on Advances in Steel Structures,
035010. Hong Kong Institute of Engineers, Lisbon, Portugal.
Stillmaker, K., Kanvinde, A. M., and Galasso, C. (2016). “Fracture me- Wen, H., and Mahmoud, H. (2016a). “New model for ductile fracture of
chanics based design of column splices with partial joint penetration metal alloys. I: Monotonic loading.” J. Eng. Mech., 10.1061/(ASCE)
welds.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X.0001380, EM.1943-7889.0001009, 04015088.
04015115. Wen, H., and Mahmoud, H. (2016b). “New model for ductile fracture of
Suresh, S. (1998). Fatigue of materials, 2nd Ed., Cambridge University metal alloys. II: Reverse loading.” J. Eng. Mech., 10.1061/(ASCE)EM
Press, Cambridge, MA. .1943-7889.0001010, 04015089.
Wallin, K. (1998). “Master curve analysis of ductile to brittle transition Xue, L. (2007). “Damage accumulation and fracture initiation in uncracked
region fracture toughness round robin data.” Technical Research Center ductile solids subject to triaxial loading.” Int. J. Solids Struct., 44(16),
of Finland, Espoo, Finland. 5163–5181.
Downloaded from ascelibrary.org by University of Waterloo on 11/12/16. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 03116001-15 J. Struct. Eng.

J. Struct. Eng., 03116001

You might also like