Sei sulla pagina 1di 45

Reprinted from JAI, Vol. 9, No.

1
doi:10.1520/JAI103940
Available online at www.astm.org/JAI

R. Sunder1

Unraveling the Science of Variable Amplitude


Fatigue
ABSTRACT: Conventional methods to estimate variable-amplitude fatigue
life revolve either around cumulative damage analysis using the local stressstrain approach, or, around one of the crack growth load interaction models.
Despite advances in modeling the mechanics of fatigue, none of these methods can faithfully reproduce the near-threshold variable amplitude fatigue
response that determines the durability of machines and structures primarily
because they fail to model the science behind the residual stress effect. Residual stress effects have a strong bearing on metal fatigue and owe their
inuence to the moderation of crack-tip surface chemistry and surface
physics. This demands the treatment of threshold stress intensity as a variable, sensitive to load history. The correct estimation of crack closure is also
crucial to determining the variable amplitude fatigue response and demands
assessment of the cyclic plastic zone stress-strain response.
KEYWORDS: fatigue crack growth, variable-amplitude loading, crack
closure, residual stress

Introduction
Many complex phenomena of engineering signicance including heat transfer,
stress/strain distribution in materials and built-up structures, their dynamic
response, and even uid ow have been understood to a point where analytical

Manuscript received May 2, 2011; accepted for publication November 1, 2011; published
online December 2011.
1
BiSS Research, 41A 1A Cross, AECS 2nd Stage, Bangalore 560094, India, e-mail:
rs@biss.in
Presented at the 11th ASTM/ESIS Symposium on Fatigue and Fracture Mechanics, Anaheim, CA, USA, May 17-20, 2011. Submitted for publication in ASTM STP.
Cite as: Sunder, R., Unraveling the Science of Variable Amplitude Fatigue, J. ASTM
Intl., Vol. 9, No. 1. doi:10.1520/JAI103940.
C 2012 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West
Copyright V
Conshohocken, PA 19428-2959.
20

SUNDER, doi:10.1520/JAI103940 21

and numerical modeling, practically from rst principles, can simulate the
actual process with amazing consistency. In stark contrast, the science of metal
fatigue has remained largely empirical even after 150 years of intense study. Incredible improvements have been effected in the safety and useful life of such
heavily stressed transportation vehicles such as aircraft and automobiles.
These were made possible to a large extent by advances in analytical techniques
related to stress-strain distribution in materials and structures under both
static and dynamic conditions, and in the area of materials engineering. The
quality of computer-aided design through solid modeling and nite element
analysis permits even less experienced engineers to ensure a uniform distribution of stresses and avoid localized stress concentration, so that adequate
safety factors can be provided without substantially increasing weight. Finally,
fracture mechanics combined with improvements in non-destructive evaluation (NDE) allows on-condition maintenance, whereby structures and
machines can be periodically inspected and repaired or retired only if necessaryif NDE does not reveal a defect, the structure must be good till the next
inspection.
A brief review of progress in understanding metal fatigue is made below in
an attempt to explain its enigmatic nature. This is followed by a description of
two major operative mechanisms that control variable-amplitude fatigue, crack
closure, and residual stress. The implications of the synergy of the two independent phenomena are discussed. The paper concludes with a description of
new avenues for research that follow from the discovery of the science behind
the residual stress effect and improved crack closure measurement.
Metal FatigueA Chronological Brief
Crucial Early ObservationRailway engineers in the early 19th century
were shocked to discover that wagon axles made from high quality ductile
steel could inexplicably break like glass, even though operating stress levels
were far less than the tested static strength of these superior quality steels.
Thus, the same material would show a brous (ductile) fracture when it fails
statically and a crystalline (brittle) one when it fails under very long term
repeated loading of low magnitude [1]. This gave birth to the speculation
(theory at the time), that cyclic loading can induce metallurgical transformations even at ambient temperature, forcing local brittle failure along crystallographic planes. Steam from the locomotive owing past axles was cited as one
possibility [2]. The present study proposes, in part, to show that while such
conclusions may seem delusive, the factual signicance of the crystalline
appearance of high cycle fatigue fractures appears to have been overlooked for
too long.
Significance of Cyclic LoadingWohlers experiments in the midnineteenth century opened up metal fatigue to engineering applications [3].
He established the concept of the S-N curve that relates fatigue life to the amplitude of cyclic loading. By performing tests at higher stress amplitudes,
Wohler showed that fatigue fractures could retain the brous appearance

22 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

associated with static fracture.2 He also established the idea of a fatigue limit
and its relationship with mean stress. In so doing, Wohler put in place the idea
of fatigue being sensitive to both the amplitude and mean level of cyclic loading and also the machinery of empirical correlation that continues to serve as
the foundation of fatigue analyses. The signicance of Wohlers work must be
judged against the background of prevailing speculative interpretations of the
time along with the backdrop of the Industrial Revolution. Scientic advance
of the discipline came much later through its association with cyclic slip, as
summarised in Fig. 1. This perception served as virtual blinders, clouding for
more than a century, a pertinent but inconvenient question: if fatigue is indeed
driven by cyclic slip, why is fatigue life and particularly, fatigue limit, so sensitive to mean stress?3 The link between cyclic plastic strain, reversed slip, and
dislocation dynamics appeared to hold much more promise given the nebulous nature of the mean stress effect. Additionally, with the subsequent discovery of crack closure (to which we will return), the mean stress effect also
appears to have been treated as effectively closed.
Cumulative Damage and Service Load EnvironmentService loading typically involves a mix of cycles of varying magnitude and asymmetry, with the
largest load occurring extremely rarely in actual usage, if at all.4 Merely ensuring that stresses due to the largest expected load do not exceed the fatigue limit
is an impractically safe design proposition except, perhaps, in civil structures.
The Miner Rule5 introduced in the early 20th century attempts to resolve this
problem by suggesting that the remaining life in a given variable-amplitude
load history undergoes a continuous cycle-by-cycle fractional decrement
expressed as the inverse of total fatigue life after each load cycle [13]. Thus, for
any given arbitrary load sequence, failure is associated with the sum of
2)

In commenting on Wohlers collection of laboratory fatigue fractures displayed at the


Paris Exhibition in 1867, Anon. prophetically observed M. Wohlers modest exhibition
may have been overlooked by ninety nine out of a hundred professional visitors to the Exhibition, yet we believe ourselves justied in saying that his scientic and patient experiments will be referred to long after the majority of those things which have drawn a
shower of medals and ribbons upon themselves at present will be dismissed and forgotten [4]. Indeed, in terms of value, Wohlers lifetime effort appears formidable even
given todays experimental resources. Just consolidating the results of his fatigue experiments under a vast variety of conditions involving axial, shear, and torsional loading
would constitute a meaningful research effort.
3)
Particularly considering that cyclic slip is mean stress insensitive! From the published
literature, only Mansons expression of hope that a meaningful rationale for the meanstress effect would be a noteworthy achievement over the coming 25 years [5] appears to
suggest awareness of the enigma surrounding an important but unresolved phenomenon.
4)
Examples are the occasional potholes for automobiles and turbulent weather for aircraft. Careless driving over deep potholes and a ight straight into a storm may serve as
extreme design considerations.
5)
Though it is known this way, actually, the rule was proposed some 20 years earlier by
Palmgren in Europe.

SUNDER, doi:10.1520/JAI103940 23

FIG. 1A brief on metal fatigue. (a) Typical fatigue test results obtained in Wohlers
time [7] shown as tables of max applied stress (fully reversed in tension and compression by rotation-bending) versus cycles to failure. (b) Test results of Wohler and Baushinger for different steels showing that the fatigue limit is mean stress sensitive [8].
Many decades later, these came to be better known as the Goodman diagram [9]. (c) A
new understanding of fatigue emerged with the association of yield with dislocation
movement. Motts analog between slip and the ease of moving a fold in a carpet and
[10,11] helps explain the formation of persistent slip bands (PSBs) (d) [12]. This, in
turn, readily explains why fatigue life is controlled by the plastic strain range (e). (f)
Cycles A, B, and C, being identical in magnitude, will cause the same extent of reversed
slip or cyclic plastic strain. They ought to result in the same fatigue life, but do not, as
shown by Wohler and Bauschinger in (b). This has been an enduring enigma surrounding metal fatigue.

24 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

cumulative fractional damage from successive load cycles attaining unity. The
idea of cumulative damage is purely notional, carries no scientic rationale,
and is not associated with any entity that could be monitored in real time.
Nevertheless, it held out the promise of practical application in designing for
desired nite life, such as the warranty period for non-safety critical engineered
products. Any such optimism was soon dashed by Gassners experiments under
multi-step programmed block loading [14]. He established that the actual damage sum at failure can uctuate wildly, depending on the mix of programmed
loads, i.e., that fatigue damage is not linearly cumulative. In the tumultuous
years preceding WWII, Gassner proceeded to develop empirical procedures
involving testing under a simulated service environment, in order to obtain fatigue life curves valid for a given material, component, joint, or even structural
assembly, subject to the statistical equivalent of a given service load history.
Thus, while Gassners effort did nally come up with an engineering solution, it
did so without casting any light on why metal fatigue is so sensitive to load
sequence. Continued emphasis on laboratory testing under a simulated service
environment underscores the signicance of load sequence sensitivity. In the
meantime, some four decades after Gassner experiments, the rst analytical basis to account for it emerged in the form of the local stress-strain (LSS)
approach.
Local Stress-Strain Approach
Figure 2 summarises the LSS approach that is based on the principle that notch
fatigue response will be the same as smooth specimen fatigue response to the
simulated notch root stress-strain response. Due to the hysteretic6 nature of the
notch root inelastic stress-strain response, local tensile yield during an overload
will cause a downward shift in the local stress response to subsequent elastic
loading. Assuming that fatigue is a localized phenomenon, it would follow that
accounting for sequence sensitivity of metal fatigue hinges on the capability to
simulate the notch root inelastic response and then translate that response into
local stress-strain cycles, identiable for the purpose of a cumulative fatigue
damage estimate after correcting for sequence sensitive local mean stress. The
LSS approach is built around several important advances in applied mechanics.
Neuber came up with a simple equation that relates remote elastic loading to
local inelastic stress-strain at a notch root subject to shear [15]. This was
assumed to be extendable to the axial stress-strain response. A simultaneous solution of Neubers equation with the Ramberg-Osgood equation [16] yields the
local inelastic stress-strain response to a given applied load. In the late 1960s,

6)

Deviation from linear response due to yield imposes hysteresis upon load reversal. As a
consequence, local stress and strain at any point of time need not be uniquely related to
applied load. They will become sensitive to load history and also to the direction of the
load change. Quite simply, hysteresis induces either reduced local stress at the cost of
increased local strain, or vice versa.

SUNDER, doi:10.1520/JAI103940 25

FIG. 2Fatigue damage caused by the two sequences shown in (a) would appear similar, gauging from the smooth specimen elastic response in (b). However, if the two
sequences are applied on a notch root seeing the local inelastic response as in (c), the
local mean stress in cycles B and E will be dissimilar. Thus, if Miners Rule appeared to
apply to (b), it needs to be adapted to (c) by accounting for load sequence sensitivity of
the notch root mean stress. (d) and (e) Local Stress Strain (LSS) approach serves as the
foundation of contemporary industrial fatigue design. It incorporates (d) Neuber conversion based on the Masing model of material stress-strain memory [17,18], (e) Rainow cycle counting to determine closed fatigue cycles, (f) damage estimates using
strain-life data and Miners Rule. In practice, case (b) also exhibits load sequence sensitivity, rendering the LSS approach questionable.

26 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

Wetzel [17] employed the emerging power of digital computers to combine a


linearized Masing model representing material memory effect in stress-strain
response7 with the Neuber equation into a numerical model, capable of realistically simulating the notch root cyclic inelastic response to an arbitrary applied
load sequence. This made it possible, for the rst time, to visualize the effect of
load history in inducing changes to notch root residual stress and thereby
account for its effect on fatigue damage [18]. Around the same time, Endo [19]
came up with the Rainow cycle counting technique to identify closed fatigue
cycles from an arbitrary random sequence of peaks and valleys, which is typical
of the service load environment.8
The early 1970s nally saw the emergence of a numerical apparatus built
around the Neuber conversion, the Masing model, Rainow, and cumulative
damage estimates to calculate notch fatigue life. A timely addition to fatigue
technology in the 1960s were computer controlled servo-hydraulic testing
machines. They permitted the determination of cyclic stress-strain characteristics for use in modeling the material response. They also permitted testing
under both total strain and plastic strain control, so as to obtain strain-life data
under highly controlled conditions.
The LSS apparatus was amenable to variations in terms of equations to calculate damage and correct it for sequence-sensitive mean stress. It was also
open to sophistication in terms of accounting for strain hardening and softening, stress relaxation, and creep-fatigue interaction.9 Continuous advancement
in computing power combined with its integration with nite element analyses
now permit the digital simulation of the cyclic stress-strain response at hot
spots in a structure for design optimization and durability assurance. Such software packages form the backbone of contemporary industrial fatigue design.
Even so, fatigue critical components are released into the market only after rst
testing their durability and structural integrity in the laboratory under simulated service conditions.
The continued need for component-level testing may not merely be a measure of insurance against the unexpected, but an acknowledgment of the

7)

The stress-strain curve of a material can be divided into a number of linear segments.
Metals have this amazing property to remember exactly how much they have deformed
along each linear segment and, therefore, how much more they can afford to deform
along the same segment. Thus, having exhausted one, their response will move on along
the next segment and so on. By simulating this response, one can digitally simulate a
tension-compression stress-strain response in a manner that will be remarkably similar
to that of real materials.
8)
The salient feature of Rainow is its physical consistency. Rainow counted cycles will
always correspond to fully closed stress-strain hysteresis loops required to estimate cumulative fatigue damage. Previous cycle counting techniques did not carry a physical
basis.
9)
This opened the opportunity for the research community to come up with fairly diverse
ways of computing damage through a variety of corrections employed to suit observed
empirical results, while essentially using the same technique to compute inputs in the
form of local stress and strain.

SUNDER, doi:10.1520/JAI103940 27

unknown with regard to variable amplitude fatigue. This possibility is underscored by a serious shortcoming of the LSS approach, as illustrated in Fig. 3.
For all its sophistication, even the most modern machinery of notch fatigue simulation cannot explain sequence sensitivity under a fully elastic notch root
response. Designers strive to ensure that local stresses never exceed yield. This
effectively implies that if machines and structures respond in real life the way
they do in simulation, there will be no local inelasticity.10 Experience shows
however, that while the notch root stress-strain response in real life may remain
elastic and therefore, sequence insensitive, sequence effects, in fact, become
more signicant with reducing overall stress level. This serious anomaly appears
to have remained largely unnoticed in the shadow of the elegance of numerical
simulation.
Limitations of the LSS approach should not come as a surprise. In scientic
terms, advances over what Wohler had originally conceived some 150 years earlier were restricted to the newfound ability to accurately determine the local
stress strain response at fatigue critical locations. Note that local stress and
strain amplitude is load sequence independent.11 Their estimation does not
actually require the elaborate cycle-by-cycle numerical simulation provided by
state-of-the-art software. The only reason for resorting to cycle-by-cycle simulation is to determine sequence sensitive local mean stress. If, indeed, this sensitivity disappears under a fully elastic response, there must be other reasons for
metal fatigue being load sequence sensitive. The LSS approach elegantly handles the mechanics of the notch root response, however. it fails to address the
science behind how such mechanics induce fatigue damage and, particularly,
why such damage may be sensitive to mean stress. Viewing fatigue as largely a
process of crack growth opens the possibility of resolving this problem (Fig. 4).
The impressive analytical machinery upon which the LSS approach is based
may indeed provide an accurate picture of the sequence-sensitive notch root
cyclic inelastic stress strain and cycle-by-cycle variation in residual stress under
service loading. However, fatigue crack growth consumes the bulk of total fatigue life and unlike a notch root, the crack tip will, by denition, always see an
inelastic cyclic response. Thus, once a crack appears, sequence effects will not
only continue to prevail under the elastic notch root response, but may even
become dominant, given the nature of near-threshold crack growth sensitivity
to overloads. Obviously, one cannot hope to harmonize variable amplitude fatigue test results obtained using the LSS and fracture mechanics approaches as
shown in Figs. 4(c) and 4(d).

10)

Note that cyclic inelasticity demands the exceedance of twice the yield stress, rendering it even more improbable in durable designs. However, even such designs often ultimately fail in fatigue, suggesting that in real-life cracks can form and grow even in the
event of totally elastic notch root response.
11)
Local stress and strain amplitude are uniquely related to applied stress amplitude by
the Neuber and Ramberg-Osgood equations, stress concentration factor, Youngs modulus, the strain hardening exponent, and cyclic strength coefcient. Applied mean stress
and mean strain do not gure in the relationship.

28 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

FIG. 3(a) Computed fatigue life versus local elastic design stress using the LSS
approach for an airframe structural Al-alloy under typical ghter aircraft (FALSTAFF
[20]) and transport aircraft (TWIST [21]) load spectra [22]. The shaded area is the estimated potential variation due to load sequence rearrangement. Note that curves for
both spectra merge into a single line below twice the yield stress (800 MPa), when cyclic
slip turns negligible. (b) Schematic notch root response for symmetric load spectrum,
and (c) response for asymmetric spectra such as FALSTAFF and TWIST. Even assuming twice the yield strain at the highest load, only symmetric spectra such as rotating
parts seeing fully reversed loading are likely to experience cyclic inelastic conditions.
Others, as in (c) will not see cyclic inelasticity and, according to the LSS approach,
should not exhibit sequence sensitivity. However,in practice they do, and do so to a signicant extent, undermining the credibility of the LSS approach. Sequence effects obviously have to do with the nature of fatigue crack growth. Crack tip response will always
be sequence sensitive because the crack tip will always see a cyclic inelastic response.

SUNDER, doi:10.1520/JAI103940 29

FIG. 4(a) Fatigue as a crack growth process. Advances in non-destructive inspection


technology are likely to increase demands on the ability to model the growth of smaller
cracks at lower growth rates. (b) Fractograph of natural crack formation and growth
under 3-step programmed loading in an Al-alloy out of an inclusion seen at bottom left.
Each band corresponds to 2000 cycles and is indicative of the reproducibility of the fatigue crack growth process even at small crack size and low growth rates [25]. (a) and
(b) Are suggestive of fatigue as a crack growth process, sensitive to crack tip cyclic
response, rather than of cumulative damage at the notch root. (c) and (d) Range and
damage exceedance (RDE) curves computed for Al-alloy L73/2014-T6 under FALSTAFF
and TWIST load spectra [26]. 1Rainow counted cycle range; 2damage contribution calculated using the LSS approach at 800 MPa (see Fig. 3(a)), and contribution to
fatigue crack extension for a small crack [3] and long crack [4]. Note that in FALSTAFF, just 10% of the cycles (the largest) contribute in excess of 90% of the damage.
This explains why the MiniFALSTAFF and FALSTAFF spectra yield similar results. On
the contrary, in the case of the TWIST spectrum, the LSS and fracture mechanics
approach provide contradictory results, with the former wrongly indicating that just
some 2% of the cycles contribute all the damage, while in actual experience, the smaller
cycles control damage. As shown by curves 3 and 4, when small cycles determine crack
growth, load interaction effects gain in importance. This underscores the signicance of
the near-threshold behaviour and its potential load sequence sensitivity.

30 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

Modern fatigue critical structures including most airframes are periodically


inspected for cracks. If no cracks are observed, the structure is released for further service until the next scheduled inspection. This implies indenite usage,
provided cracks, if detected, are immediately repaired, or the part is replaced.
The cost of repair will eventually determine retirement for cause [23]. The
cost of inspection, along with its periodicity, will determine the overall economics of operation. In this scheme, the enforced periodicity of inspection is determined by the quality and reliability of non-destructive inspection (NDI), which
needs to be matched by the ability to correctly estimate the residual life of the
structure with such a crack. Obviously, neither the actual initial defect size
(assuming it is smaller than NDI-detectable size) nor the ability to correctly
model very early growth carry value in a condition monitoring scheme.
From the overall standpoint of durability assessment, understanding fatigue crack growth response below NDI-detectable crack size becomes valuable
in the event there is a demand for an extended period of service before rst inspection. It assumes even more importance when the component is not subject to
inspection. Additionally, it certainly offers the promise of just doing away altogether with the obsolete concept of cumulative fatigue damage. The potential
for doing so is supported by the highly reproducible growth bands in Fig. 4(b)
even at incredibly small crack sizes.
As a rule, the quality of life estimate is inversely proportional to life [24].
Assuming the bulk of that life is exhausted by crack growth, the study of near
threshold variable-amplitude crack growth becomes extremely important.
Indeed, the potential for the advancement and application of fracture mechanics in structural design over the last four decades has largely overshadowed
opportunities presented by the LSS approach.
Fracture Mechanics Approach
With the birth of linear elastic fracture mechanics, the stress intensity factor K
became available, that serves several important purposes. Here, K is, in effect, a
similarity criterion, to which both residual strength and fatigue crack kinetics
can be related (see Fig. 5). Paris showed that the fatigue crack growth rate da/
dN correlates with the cyclic stress intensity range DK [27]. This was a turning
point in the advancement of fatigue research. In contrast to a notional parameter called cumulative damage, a quantiable parameter in the form of crack size
was now available to characterize damage. Further, K permits the unication of
experimental data for a given material, irrespective of cracked body geometry,
crack size, shape, and applied load level. In effect, K is to a cracked body what
stress is to a smooth uniform section specimen. Using K, experimental crack
growth data obtained on simple laboratory coupons could be readily extrapolated to structural components of engineering interest.
Crack Growth Load Interaction Models
The 1960s saw much progress in unraveling the mystery behind the load
sequence effect researched forty years earlier by Gassner that had debunked the

SUNDER, doi:10.1520/JAI103940 31

FIG. 5Stress intensity factor K as a similarity criterion for fatigue crack growth. (a)
Stress intensity for crack subject to uniform remote stress [1] increases with crack size
which is the inverse of the case of rivet (point) load [2]. Correspondingly, the growth
rate, da/dN will also vary differently with crack size. Yet, as shown in (b), da/dN for the
two cases will fall into a single scatter band when plotted against the stress intensity
range [28]. Experience shows, however, that the relationship (b) combined with K are
not sufcient similarity criteria for engineering applications. Consider the schematic of
the loads in (c) on a transport aircraft at Atake-off and climb, Bcruise, and C
descent and landing (load level on a transport liner gradually drops due to mass reduction from fuel consumption). Crack growth curves will vary as shown in (d), depending
on the mere rearrangement of loads [29]. Cycles covering a few thousand ights and rearranged to form a Hi-Lo programmed sequence will yield a crack growth life about
four times greater than if applied as is. This is attributed to load interaction mechanisms including crack closure, residual stress, and crack front incompatibility.

32 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

Palmgren-Miner Rule. The advent of precision servo-hydraulics based test systems allowed systematic experiments on variable amplitude fatigue crack
growth. These permitted the study of crack growth rate transients after overloads and underloads superposed on baseline constant amplitude loading.
Experiments came up with the astonishing nding that applying a tensile overload, in fact, ends up retarding further crack growth even if the crack would have
substantially incremented during the overload. It was also found that compressive overloads (inappropriately called underloads) could, in effect, erase the
retarding effect of a previous overload. These observations revealed that under
variable amplitude loading, the order in which different loads are applied inuences the rate of crack advance in a manner that could not be readily explained
by considerations of solid mechanics (see Fig. 5(c) and 5(d)). Clearly, the material at the crack tip appeared to remember what previously transpired in a
manner that affected its subsequent fatigue resistance. The search was on for
load interaction mechanisms that may be responsible for sequence effects.
Wheeler [30] and Willenborg [31] came up with empirical models on the
consideration that the tensile monotonic plastic zone ahead of the crack tip will
act as a wedge squeezed by the elastic matrix to create a zone of compressive residual stresses at the crack tip (see Fig. 6). If an overload is applied, this plastic
zone will increase in size as a square function of the overload ratio, leading to a
substantial increase in the near-tip compressive stress. To account for this
effect, Wheeler introduced a transient retardation factor as a power function of
the ratio of remaining crack extension in the overload plastic zone to the size of
this zone with constants empirically selected to approximate experimental
observations. Willenborg interpreted the same effect in terms of a reduced
effective stress ratio due to increased compressive residual stress, also with a
transient function to t real observations. This model relies on Walkers equation correcting the growth rate for the stress ratio [32].
If, in the 1970s, the LSS approach was already incorporated into commercially available industrial software for fatigue design, the Wheeler and Willenborg
models were also brought into the market for the safe-life and fail-safe design of
aircraft structures and later, into the nuclear, piping, energy, railroad, automotive, and other industries. Forty yearslater, software built around these models
continues to dominate industrial fatigue design. Even so, safety critical designs
are invariably tested in the laboratory under simulated service conditions.
Fatigue Crack Closure
Just when it seemed that the Wheeler and Willenborg models appeared to hold
promise in application, if not in scientic conviction, Elbers [33] discovery of
crack closure (Fig. 6(e)) nally developed a mechanism that actually makes scientic sense and can be analytically modeled using fracture mechanics concepts. Newman [34], de Koning [35], and others came up with numerical
models of how the plastically stretched wake behind the crack tip effectively
closes even under tensile load. This was a milestone in the analytical simulation
of the mean stress effect in metal fatigue. What is more, the new approach was
able to simulate, with reasonable conviction, the consequences of tensile and

SUNDER, doi:10.1520/JAI103940 33

FIG. 6General scheme of load interaction models in current use. The action of a tensile overload (a) is described in (b)-(d). A is the monotonic plastic zone from baseline
loading and B, the cyclic plastic zone. C is the overload plastic zone and D, the cyclic
plastic zone due to overload, that vanishes upon the next tensile cycle. E is the crack
wake zone squeezed into bearing by the surrounding stretched material from the plastic
zone. (b) Indicates the crack tip picture upon the application of tensile overload. (c)
Shows the picture when the crack is almost through the overload plastic zone, and (d)
indicates crack tip growing through overload stretched wake. (e) Crack tip response to
load sequence 1-5, shown in the inset. Laser interferometry [36] estimates over
0.15 mm gauge length after deducting the elastic response. The loop shape unambiguously underscores the portion of load cycle when the crack was open. Also note that closure is cycle sequence insensitive (2,4 and 1,5 indicate similar closure level). This is
proof that closure is insensitive to the cyclic plastic zone response (to crack-tip residual
stress). According to both the Wheeler and Willenborg models, compressive stresses in
the overload plastic zone will retard crack growth until the baseline monotonic plastic
zone begins to exit the overload plastic zone, as in (c). Using Elbers closure model, retarded growth will persist for some distance beyond the overload plastic zone (d). Neither the Wheeler/Willenborg nor the closure models can explain the possible differences
in crack extension between cycles 2,4 and 1,5. In fact, the rst two actually model closure, even if they may profess to model the residual stress effect!

34 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

compressive overloads.12 Clinching evidence appeared by way of the ability of the


closure model to explain accelerated crack growth after a step-wise increase in
load and the ner aspect of delayed retardation after an overload. There was no
way for the Wheeler and Willenborg models to explain such behaviour. Closure
considerations make it obvious and simple. When a tensile overload is applied, it
takes some crack growth for the overload induced wake with extra stretch to take
effect. Therefore, retardation is not immediate. In fact the crack may even momentarily accelerate because the the overload itself opens up the crack, causing a
reduction in closure stress. However, when a compressive overload is applied the
consequent reduction in closure stress is immediate. Crack closure based models
were thus able to simulate, through mechanics based computations, many seemingly complex load sequence effects that had hitherto appeared inexplicable. In
crack closure, a scientic explanation at long last seemed available for the effect
of both mean stress (stress ratio) and residual stress. All other load interaction
mechanisms appeared either insignicant, were perhaps manifested through closure, or, an outright gment of imagination. Or so it seemed.
The 1970s and 1980s saw the publication of over a thousand papers related to
crack closure. The bandwagon soon became an overcrowded train, with individual
coaches representing the variety of sources of crack closure. As it were, Elbers discovery was merely of plasticity induced closure. To this were added oxide-induced
closure, roughness-induced closure, and asperity-induced closure. It was then suggested that closure is but one shielding mechanism for a fatigue crack, with the further division of shielding into extrinsic and intrinsic. Therefore, closure was now
bracketed with crack tip shielding mechanisms such as uncracked bres in the crack
wake, or, higher stiffness bres ahead of it. As a consequence, if everything seemed
simple and straightforward as illustrated by Elbers early work, a much more complex and confusing picture seemed to emerge from subsequent research.
The cause of closure has not been helped by an unfortunate aspect of its measurement. Unlike parameters that can be directly measured, such as dimensions or
weight, or at least by an easy to dene and strictly reproducible process such as modulus of elasticity, yield stress, or ultimate stress, crack closure measurement carries a
heavy measure of interpretation. An annexure to ASTM E647 with a recommended
practice for closure measurement is a good example of a technique that delivers
measurements of little practical value. Remote measurement of crack opening displacement representing contact response integrated way beyond intervals actually
affecting closure carries only a remote chance of correlation with an actual value.13
12)

The Wheeler and Willenborg models could not account for the effect of compressive
overloads.
13)
Closure induces a certain wedge opening stress intensity to compensate for the applied
stress falling below closure stress. The contribution to the stress intensity of a point force in
the crack wake will be inversely proportional to its distance from the crack-tip. Assuming compressive yield stress upon wake contact, the depth of relevance to closure is of the order of a
monotonic plastic zone size. Displacement measurements made remote from this zone of
inuence cannot be expected to sense the crack tip response with the desired sensitivity.
Indeed, there are no published data showing credible closure measurements under variable
amplitude loading.

SUNDER, doi:10.1520/JAI103940 35

The issue is further complicated by difculties in mechanism isolation to eliminate


ambiguity in the interpretation of the results. For example, would it be fair to attribute retarded crack growth to roughness induced closure, when the very occurrence
of roughness may have also reduced the intensity of the crack tip stress eld by a
ragged crack front and possible multiple plane separation?
The technique in Fig. 6(e) involving near-tip laser indentation interferometry14 [36] and fractography using the Closure Block15 [37] are exceptions that
deliver reproducible and scientically defendable results. Unfortunately, these
are not amenable to easy implementation in routine engineering laboratory
measurements.
Load Sequence Sensitivity of Individual Crack Extension Mechanisms
We now proceed to analyse different stages of fatigue crack growth associated
mechanisms and how they may be affected by the variable-amplitude environment. Measurable fatigue crack growth rates range from less than atomic spacing, right up to 1 mm/cycle, a potential variation of at least eight orders of
magnitude (Fig. 4(a)). There are not many phenomena of engineering relevance,
with such a wide swing in kinetics. Crack extension itself occurs in an environment of several competing mechanisms, with individual mechanisms dominating selected intervals of growth rate. Add to this the different ways in which
ambient conditions can affect individual mechanisms. It would, therefore,
come as a surprise if any single crack extension mechanism can describe the
process. Even more surprising would be a single load interaction model coming
up with consistent estimates of variable-amplitude crack growth rates. For
clarity, we broadly divide crack kinetics into three distinct ranges of the crack
growth rate and proceed to examine how the dominant crack extension mechanism in each range responds to variable-amplitude loading. Before doing so, we
dene a basic assumption that is required to distinguish fracture mechanics
based analysis from cumulative damage concepts.
History Effect on Crack ExtensionConsider crack extension in identical
cycles A, B, and C shown in Fig. 7(a) with different loading histories. Case (a)
involves constant amplitude loading. Cases (b) and (c) involve prior cycling at
increased loading amplitude, causing greater near-tip cyclic slip. Based on cumulative fatigue damage considerations, one should expect crack extension C to
exceed B and for both to exceed A due to greater prior damage, However,
there appears to be absolutely no empirical evidence to suggest such a possibility!16 Fracture mechanics based models of variable-amplitude fatigue, in fact,
14)

With a working gage length of the same order as the plastic zone size, this technique is
sensitive to the inelastic stress-strain response within the cyclic plastic zone as seen in
Fig. 6(e).
15)
The technique proceeds on the premise that given constant Kmax, there is no other explanation for equal striation spacing under varying Kmin other than equal DKeff.
16)
Not necessarily because such a possibility does not exist, but rather, because of the limitations in experimental techniques to address the question in quantiable terms.

36 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

FIG. 7(a) According to the cumulative damage concept, identical load cycles A, B,
and C may extend the crack differently because of the different load history preceding
each of them. In contrast, all crack growth models ignore the possibility of damage to
material ahead of the crack tip. This understanding is central to analytical modeling of
load history effects. (b) The three growth rate regimes and their associated fractures for
an Al-alloy. Crack extension in a cycle under variable amplitude loading may fall into
any of these three regimes, depending on its magnitude. (c) During the rising half cycle
shown in the inset, the crack will rst extend by brittle micro-fracture (BMF) over a nite number of atomic layers embrittled by instantaneous surface diffusion (ii), and
then switch to shear extension (iii), suggesting striation formation by the mode change
(iv) [39]. Any further increase in load beyond 2 may induce a disproportionately higher
quasi-static crack extension. This explains why striations marking individual cycles are
seen only over a very narrow range of growth rate.

SUNDER, doi:10.1520/JAI103940 37

simply ignore it. They assume that the crack extension in the next cycle is driven
only by the magnitude of that cycle. The prevailing understanding of crack
growth load interaction effects is also based exclusively on variables that control
crack kinetics in the next load cycle. It ignores any prior slip-reversal damage
to the crack tip. In the absence of compelling arguments to the contrary, we
shall ignore any prior damage and its effect in considering dominant crack
extension mechanisms and how they respond to variable amplitude loading. In
doing so, we make an important assumption that the fatigue crack can extend
under each load cycle.17
Dominant Crack Extension MechanismsAt the commencement of the rising
half of a new load cycle, the dominant crack extension mechanism is still an
unknown. Crack extension will commence by a yet to be dened mechanism once
the load excursion exceeds a certain threshold value. It will soon transform to striation mode as the stress intensity falls into the Paris regime and then proceeds to
extend through local quasi-static fracture in the event K approaches critical values
(see Fig. 7(b)). Each of these three stages occupies a nite but overlapping interval
of crack growth rates, with the rst transition occurring around 104 mm/cycle
and the second one depending largely on the stress ratio, around 102 mm/cycle.
With the increasing stress ratio, this last transition will progressively move into
lower growth rates because of the onset of quasi-static fracture leading to a shortened Paris interval. Note that the different stages in crack growth are associated
with the change in growth rate over several orders of magnitude. Higher order
growth rates will necessarily be associated with a mix of mechanisms18 (see
Fig. 7(c)), though the last mechanism to switch-in would emerge as the dominant
one by virtue of its disproportionately large contribution to crack extension.
The above rationale suggests that in variable amplitude fatigue, a variety of
crack extension mechanisms will continuously leave an imprint on the fracture
surface and their mix will depend on the load spectrum. A corresponding mix of
load interaction mechanisms may also continuously prevail. We now proceed to
consider in greater detail, individual crack extension mechanisms and how
each one may be sequence sensitive. In doing so, less signicant load interaction
mechanisms such as crack-tip blunting/resharpening, history-induced phase
transformations, and other such effects whose inuence cannot be deemed decisive or quantiable are ignored.
High-End Growth Rates
The crack tip will see critical conditions associated with catastrophic fracture
when K approaches Kc associated with static fracture. Such local failure is
17)
Crack growth rates less than atomic spacing are readily explained by the possibility of
local crack extension occurring at different points on the crack front at different times
[38].
18)
After all, the crack tip at the commencement of rising load half-cycle, does not yet
know the extent to which it will be loaded. It will switch sequentially to the mechanism
of least resistance to crack extension corresponding to the instantaneous load increment.

38 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

attributed to quasi-static rupture of the material directly ahead of the crack tip.
If the material is inherently brittle, it will simply cleave locally. If it is ductile, as
is the case with most aerospace structural materials, at least two simultaneous
mechanisms are likely. Stable crack growth by shear can be either Mode II or
Mode III. This typically occurs at the specimen edge, where plane stress conditions promote shear ligament formation and gradually spread inward, because
ligament formation demands crack extension.19 A little deeper, and particularly
given a straight crack front, plane strain conditions associated with constraint
can prevail, leading to the buildup of hydrostatic tension20 that can result in
static rupture by microvoid coalescence (essentially, an analog of cavitation in
liquids), seen on the fracture surface as clusters of microscopic cavities, irrefutable evidence that local failure was instantaneous. Note that because conditions of constraint develop at some distance from the crack tip, crack jump or
tunneling by microvoid coalescence will invariably be accompanied by a shear
of the interim ligament at the very tip of the crack that remained under plane
stress. A third mechanism is typical of Al-alloys and the proliferation in them of
secondary particulates that are natural barriers to slip. As a consequence, if
sizeable slip is involved that covers a distance exceeding their average spacing, a
strain localization will result, leading to a shear fracture along interconnecting
planes between particulates. This leads to the appearance on the fracture surface of a disproportionately high density of particulate voids, that should not be
confused with microvoid coalescence associated with static fracture as was the
case in. An example of a mix of the two appears in Fig. 7(b) (also, see Fig. 10(b)).
Being a highly localized phenomenon, such ruptures may occur momentarily
and only at one or a few points ahead of the crack front. This, in macroscopic
terms, will show up as increasingly accelerated fatigue cracking as Kmax under
cyclic loading approaches Kc.21 One may expect that as the ratio Kmax/Kc
approaches unity, the crack growth rate will approach innity (static fracture).

19)

As a rule of thumb, the crack needs to extend over an interval of at least half the specimen thickness in order for the front to completely rotate to shear mode. Quite simply,
front rotation also demands extension.
20)
Liquids follow Pascals Law. Applying pressure at any point will result in all ends of the
constraining container seeing that pressure. This is what drives uid power technology.
Solids are different from liquids in their resistance to sliding (shear or slip), which is
innitely higher than viscosity in liquids. Therefore, when a smooth solid specimen is
pulled, it will readily transversely contract, as seen on a rubber band. However, if for
some reason such a contraction is inhibited by external or internal conditions (constraint), a hydrostatic response will result, whereby tension will be experienced in all
directions. An example of hydrostatic tension in the response of secondary particulates is
forthcoming. A stress gradient serves as a natural constraint and can result in a nearhydrostatic local response.
21)
In the presence of a substantial quasi-static crack extension, one can hear audible popins. Much lower levels of such an extension can be picked up by acoustic emission, which
often serves as a tool for on-line structural diagnostics. This is used in industry to hear
defects growing in a structure and to locate them by triangulation, much like GPS positioning systems.

SUNDER, doi:10.1520/JAI103940 39

Forman et al. introduced such a correction into the crack growth rate equation
which otherwise only carried two material constants to be determined by statistical analysis of laboratory data. The correction kicks in only at higher values of
DK, or at a very high stress ratio, where Kmax gets closer to Kc even at lower DK.
Critical conditions associated with local quasi-static crack extension require
high stress and strain levels. Since these will be tied to the top end of the local
stress-strain hysteresis loop, they may be immune to hysteretic effects and
therefore insensitive to load history. Also at these levels, crack closure has practically no role to play because the process is driven by the maximum driving
force, rather than its range. There is, however, some possibility of effects attributable to strain hardening or softening that may affect local fracture resistance
and will be stress history sensitive. Importantly, the crack-tip stress-strain
response will be extremely sensitive to local constraint. This will vary across the
thickness and will also be determined by instantaneous crack front orientation
as well as shape, that is, in effect, determined by the cumulative preceding crack
extension. Of all the load history related parameters, this one appears worthy of
analytical consideration at a high growth rate. To do so, one may treat Kc as a
crack front related parameter varying between a low of K1c associated with
plane strain and a high of Kc, associated with plane stress and therein introduce
a history sensitive component into the Forman equation to account for
sequence sensitivity of high end growth rates.
In summary, the effect on high end growth rates of the loading history may
be accounted for by correcting K and Kc for crack front shape and orientation.
Parameters such as crack closure and residual stress will have little bearing on
high-end growth rates.
Intermediate (Paris Regime) Range Growth Rates
The Paris Regime is characterized by a log linear relationship between DK and
da/dN over a range of growth rates covering the interval 104102 mm/cycle
with nonlinearity at the high end coming in due to the quasi-static component
and with the lower end overlapping with near-threshold fatigue response. The
interval is dominated by cyclic slip driven crack-tip extension, according to a variety of schemes proposed in the literature [41]. A reasonably straight crack
front is conducive to transgranular slip along preferred planes and one can
readily accept the possibility within individual grains of highly reproducible
extent of stretch and compression in successive load cycles that leave behind
striation bands with near digital precision.22
There are different ways in which a crack can extend over a load cycle in a predominantly slip dominated mode. The rst is by deformation (as opposed to
22)
Reference [42] describes an experiment that involved punching onto the fracture surface of fatigue striations representing binary code of text strings in much the same way as
information is stored on digital media. This would not be possible without precisely reproducible cycle-by-cycle fatigue crack extension at the microscopic scale and serves as a
compelling argument in favour of fractography as a dependable tool not only in failure
analysis, but also for the quantitative validation of crack growth models.

40 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

fracture), whereby the shear stretch produced in the rising half cycle cannot be
fully reversed upon unloading, resulting in a fold, as indicated by the well-known
Laird model [43]. From this, follows the unexpected conclusion that the crack
extends during unloading. The second possibility is that the crack extends by shear
fracture [44], whose extent is determined by rising load excursion exceeding a certain threshold level over which microscopic stable crack extension occurs, but not
unstable (even if localized) fracture. Reversed deformation during unloading will
essentially prepare a sharp crack for extension in the next cycle.23 The third possibility is a combination of the two, leading to a somewhat greater crack extension
considering that the crack will continue to grow during unloading as well. All three
possibilities are supported by observations of extremely well dened striations that
mark the fatigue fracture surface, though the textbook understanding is of fatigue
crack extension by deformation (slip), not shear fracture.
Assuming that the crack-tip response is controlled exclusively by the cyclic
stress-strain curve and the extent of change in stress intensity, crack extension in
this range should be insensitive to the applied stress ratio and to near-tip mean
stress (i.e., residual stress). Mean stress insensitivity is the very essence of a process
driven by slip alone. It follows that any sensitivity of intermediate range crack
growth rates to the stress ratio and to the load history may be attributed largely, if
not solely, to crack closure. An inevitable conclusion then would be that if the
Wheeler and Willenborg models indeed correctly simulate intermediate range variable amplitude behaviour, they may be merely appearing to do so by the happy
coincidence of fudged closure response. Indeed, if fatigue crack growth is predominantly slip driven, the only plausible explanation for the stress ratio and load history
effects is fatigue crack closure controlling the effective range of the stress intensity.
All three possible ways of crack extension by slip previously listed carry certain implications that go beyond insensitivity to residual stress, stress ratio, and
stress history. They imply cycle-by-cycle striation formation. They also imply
relative immunity of the Paris Regime to the environmental effect (assuming
slip is environment independent) and to cycling frequency (assuming rateinsensitivity of slip over the practical range of frequency). Sensitivity to environment and frequency increases at lower growth rates associated with thresholds
and at much higher rates associated with sustained load cracking, creep, etc.
There are two curious features of intermediate range crack growth whose
signicance appears to have remained largely unnoticed over the ve decades of
study by high resolution electron fractography. One is the surprisingly narrow
band of growth rates (usually within one or two orders ofmagnitude of variation) over which discernible striations are observed.24 The other is the surprising absence of striations in vacuum.

23)
A blunt crack tip offers multiple parallel slip planes that will contribute to cumulative
stretch by dissipating total strain. A sharp crack restricts the number of shear planes and
thereby encourages shear fracture by focusing shear strain into fewer slip planes.
24)
The resolution of electron fractography is adequate to resolve a crack extension less
than 106 mm/cycle, but one seldom sees striations at growth rate less than 104 mm/
cycle.

SUNDER, doi:10.1520/JAI103940 41

Near Threshold Fatigue Crack Response


Indeed, why are striations not discernible in vacuum? And why are we usually
unable to see striations in atmospheric fractures at growth rates below
104 mm/cycle, even if electron microscopes can resolve features one hundred
times smaller? The controversial brittle micro-fracture (BMF) model of nearthreshold crack growth25 appears to provide the answer [45] (Fig. 8). The fatigue crack tip represents an extreme stress concentrator. Associated with such
stress concentration is an extreme stress gradient that in turn induces conditions of severe near-tip constraint because the surrounding lightly stressed material does not permit local necking. This, in turn, induces conditions of
hydrostatic loading: application of tensile load normal to the crack plane causes
increasing tensile stresses in the transverse direction as well. Stress in the third
direction along the major crack axis will be somewhat relaxed, at least at the
tip, because the free crack tip surface is free to move inward into the material.
However, the diaphragm stresses stretching the crack tip surface in two directions will increase the inter-atomic distance along the loading axis while not
allowing transverse spacing to reduce. Such conditions are conducive to the
activation of surface physics (diffusion of active species into surface layers) and
surface chemistry (chemical reaction with active species), leading to accelerated
transgranular26 fatigue crack extension [46].
In a careful study on the near-threshold fatigue fracture mode of an
Al-alloy, Gangloff et al. observe that crack extension occurred along crystallographic slip planes [47]. This by itself need not imply that crack extension
occurred by slip unless it can also be shown that the fracture plane was oriented
appropriately with respect to the loading direction27 as is the case with ductile
response and striation formation in the Paris Regime. Once a surface layer has
been embrittled, it may not matter whether Mode I (tensile rupture) or II (slip)
is involved. If Mode II was indeed involved, it would lead to the formation of
shear lips and progressive rotation of the fatigue fracture plane by 45 . However, atmospheric fatigue fracture surfaces in the near-threshold regime do not
tend to develop shear lips. They remain at and normal to the loading plane.
Compelling evidence in support of the previous rationale comes by way of
fatigue fractures obtained in salt water, air, and vacuum [48,49] under identical
loading conditions. The authors attributed the delayed transition to shear mode
in salt water and air out of early Mode I cracking to the adverse effect of environment on resistance to Mode I. A crack tip stress state is determined by shearlip formation, which, in turn, is driven by dominant macroscopic mode of crack

25)

Against the general perception of metal fatigue being associated with cyclic slip (deformation), the BMF model suggests that near-threshold fatigue crack extension occurs by
fracture.
26)
This is not to be confused with the mechanism of stress corrosion cracking associated
with the intergranular short circuit diffusion of active species that essentially leads to
crack extension by grain separation.
27)
Just as delamination in composites can occur either by Mode I or Mode II.

42 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

FIG. 8The science behind the residual stress effect in metal fatigue crack growth. (a)
When an argon bubble is inserted under ruthenium monolayers, the stretched top
instantaneously attracts active species, while the compressed region at the root of the
blister repels them [46]. (b) According to the BMF theory, the same holds true at the fatigue crack tip [45]: the active species is moisture at room temperature that is repelled
from the crack tip at minimum load, 1. During the rising half-cycle, moisture molecules
are attracted by the rising stresses at the crack tip. They react with metal to form metal
oxide and hydroxide to release hydrogen that diffuses into the substrate to embrittle and
fracture the affected surface layers under rising stress. (c) The surface physics and
chemistry described in (b) will be affected by the crack tip stress history as shown by the
schematic repeat action of load sequence 1-7. (d) If closure is reduced or absent (LoSop), cycles 2-3 and 5-6 will see hysteretic crack-tip stress-strain response. Higher stress
causes more BMF at 2-3 than at 5-6. (e) However, if the crack is partially closed during
2-3 and 5-6, both cycles will see similar reduced local stress and therefore, equally retarded crack extension. (b)(e) Underscores the signicance of the cyclic plastic zone
response in controlling atmospheric sub-critical fatigue crack growth. Closure and
Wheeler/Willenborg models are incapable of explaining cycle-by-cycle hysteretic load
interaction effects in fatigue crack growth.

SUNDER, doi:10.1520/JAI103940 43

extension, rather than by applied DK28 or even the growth rate. Thus, in high
vacuum, shear lips will form earlier than in salt water or even air, even if vacuum growth rates will be much lower, given similar loading conditions. Pippan
et al. have observed that the fatigue crack stays sharp in air and turns blunt in
vacuum29 [50] but failed to draw conclusions on how this may reect on the
crack extension mode. Embrittled surface layers will also exhibit reduced elongation. As a consequence, the crack may extend by BMF before the potential
onset of slip on neighboring planes that promote blunting, or, on the same
plane, but deeper into the substrate.
In room temperature atmospheric fatigue, BMF appears to be primarily
promoted by surface diffusion of hydrogen released by the reaction of moisture
with the crack tip surface resulting in oxide and hydroxide formation. Oxidation
appears to be an unlikely factor in BMF, a conclusion prompted by the retarded
near-threshold fatigue crack growth in dry oxygen observed by Bowles [51]. In
tests on an Al-alloy, Bowles also observed that when the environment is
switched from laboratory air to dry oxygen, striations gradually disappear, leaving a surface akin to that obtained in vacuum.30 This observation also points to
the potential role of BMF in striation formation. The BMF controls the nearthreshold fatigue response that extends up to a growth rate of between 105 and
104 mm/cycle, suggesting that surface physics and chemistry do affect tens,
but perhaps not hundreds or thousands of atomic layers at the crack tip. Perhaps crack extension by the BMF (mode I) over such a distance in the course of
the rising load half cycle, when followed by subsequent crack extension either
by shear in Mode II, or, by folding of shear stretched crack tip surface, or, by a
combination of the two leaves that distinct wavy pattern one associates with
well-dened striations. Striation formation may thus require two distinctly different crack extension mechanisms to operate sequentially (as shown by the
schematic in Fig. 7(c)). If only one of them operates as in the case below the
Paris Regime (only BMF and no slip) or in high vacuum (only slip and no BMF),
no discernible contrasting topographical feature may result to mark the progress of the crack front.
Just as room temperature near threshold fatigue is closely linked with
cycle-by-cycle crack extension by the BMF of crack-tip surface layers embrittled
by surface physics and chemistry, a similar process may control elevated
28)
The ratio of plastic zone size to thickness is often treated as a reection of the stress
state. Implicit in this assumption is a at and straight crack front. In reality, a curved
(tongue shaped) crack front or one that is tilted will both promote plane stress due toligament response.
29)
Interestingly, having obtained lucid evidence about the cause (sensitivity of crack-tip
deformation to environment), the authors seem to have failed to draw the logical conclusion about its effect (sensitivity of the crack extension mode to the environment)!
30)
Their gradual rather than immediate disappearance also raises the intriguing question of hydrogen consumption. Does hydrogen get consumed by embrittlement, or does it
escape upon BMF to affect the next layer? Partial consumption can explain the momentary persistence of BMF into vacuum. It may also explain sustained accelerated internal
cracking as in gigacycle fatigue.

44 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

temperature transgranular fatigue crack growth. The latter is accelerated by the


enhanced oxidation of crack tip layers that can considerably exceed the depth of
moisture related surface diffusion by hydrogen. In both cases, crack extension
is transgranular and involves cycle-by-cycle crack tip surface activity that accelerates crack extension by comparison to vacuum fatigue response. For this reason, in both cases, the threshold stress intensity will be much less than in high
vacuum. It thus emerges, that, if near-threshold behaviour is sensitive to diffusion kinetics, threshold stress intensity ought to be controlled by the cyclic plastic zone response!
Threshold Stress Intensity
While near-threshold fatigue crack growth behaviour has long been connected
with crack-tip surface physics and with surface chemistry [52, 53], there seems
to have been a general failure to appreciate the connection between the kinetics
of surface activity and near tip hydrostatic stress, and the sensitivity of the latter
to the stress history and to stress ratio. This may be attributed to the prevailing
stereotype of the crack tip essentially seeing an elastic, ideally plastic cyclic
response, implying a local stress ratio of R 1 (zero mean stress), irrespective
of the applied stress ratio and stress history. Such an assumption may have
assisted crack-tip elasto-plastic stress-strain analyses and may also have been
appropriate for the Paris Regime with its sizeable cycle plastic zones. However,
it appears to have clouded the signicance of the hysteretic stress-strain
response within the cyclic plastic zone in moderating near-threshold diffusion
activity at room temperature and chemical reactivity at elevated temperatures.
On the contrary, the signicance of residual stress is well known and appreciated in stress corrosion cracking, as is practiced in assessing heat affected zones
in welding and crack growth in an aggressive environment. There is, however,
an extremely important distinction between stress corrosion cracking and
atmospheric near-threshold fatigue crack growth. Near threshold fatigue crack
growth is affected by near-tip stress zones that are hydrostatic and microscopic
in comparison to those considered in stress corrosion cracking.
When crack-tip cyclic slip recedes below a certain threshold, crack growth
will practically cease in vacuum. This vacuum threshold stress intensity range is
about three times greater than effective DKth in air. Crack growth in air at DK
less than vacuum DKth cannot obviously be explained on considerations of slip.
The difference between the two when raised to the power of four and above,
covers a growth rate variation in air, exceeding two orders of magnitude. The interim interval thus covers a vital segment of the sub-Paris regime atmospheric
fatigue response that may well account for the bulk of total fatigue life.31 Fatigue kinetics over this interval will obviously be determined by the ability of
31)

In observing fatigue fractures, one may be inclined to associate the bulk of the fatigue
process with the largest observable area of the fatigue fracture. However, the bulk of fatigue life may, in fact, have been consumed in early crack growth. While assessing fatigue
fractures, it may be important not to ignore that, albeit small, region covering the crack
initiation area.

SUNDER, doi:10.1520/JAI103940 45

hydrostatic stresses within the cyclic plastic zone to moderate crack-tip surface
chemistry and surface physics. These stresses are the sum of the crack-tip mean
stress associated with the current stress ratio, history sensitive, residual stress
and their hysteretic variation while the crack remains open.32
The consequences of such a crack-tip cyclic diffusion pump may be varied. An arrested crack tip will progressively lose its resistance under the persistent onslaught of diffusing active species. Hydrogen trapped in the rising load
half-cycle will not be released upon unloading. Oxidation at an elevated temperature accelerated during the rising load half-cycle will not be reversed upon
unloading. This implies that the crack front will be inclined to straighten itself
even if the crack does not uniformly extend in successive cycles. Over each cycle
that the tip does not give way, surface layers are likely to see a little more
embrittlement. At the same time, interstitial diffusion is a self-retarding process
because diffused layers represent barriers to newer and deeper diffusion. This is
why the effect in question is unlikely to signicantly inuence growth rates in
excess of 104 mm/cycle. Another measure of the effect can emerge from a comparison of Paris Regime growth rates in air and high vacuum. Their difference
is substantially less than under near-threshold conditions. Thus, while a fatigue
crack in air can grow at 105 mm/cycle, it may just remain arrested under high
vacuum given the same loading conditions.
Significance of Cyclic Plastic Zone Response in Variable Amplitude Fatigue
Near-tip residual stress is highly sensitive to load history and can vary substantially on a cycle-by-cycle basis. Figure 8(c)8(e) schematically illustrates the
cycle-by-cycle near-tip stress history for a fully open and for a partially closed
fatigue crack. Consider the repeated action of cycles 1-7. Identical embedded
cycles 2-3 and 5-6 will see the consequences of the hysteretic crack tip response
within the cyclic plastic zone. In both cases, near tip stress will be well below
that under constant amplitude loading because of the compressive stress introduced by tensile overload 4. As a consequence, cycles 2-3 and 5-6 will see retarded crack extension. However, the retardation in 5-6 will be greater because
lower local stress reduces the diffusion activity with the crack growth tending
towards vacuum response. Importantly, the difference between 2-3 and 5-6
being hysteretic can be seen on a cycle-by-cycle basis. This is possible only if the
crack is fully open during the embedded cycles as in Fig. 8(d) and with closure
stress well below the minimum stress in the two cycles. The hysteretic variation
will cease in the event of partial closure as shown in Fig. 8(e). Both cycles will
be equally retarded in this case, being rendered cycle-sequence insensitive due
to partial closure. Cycle-sequence sensitivity is a term alien to conventional
modeling based on the Wheeler, Willenborg, or Elber models.33 The signicance
32)

An important consequence of this possibility is that a partially closed crack will not see
cycle-sequence sensitivity, a feature to be addressed further in the text.
33)
Curiously, interpretation of notch root fatigue response universally proceeds on this
very understanding, and has remained unquestioned, even in the absence of any scientic
rationale for the notch root mean stress effect!

46 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

of the cyclic plastic zone and the possibility of cycle-sequence sensitivity


becomes obvious only when viewed from the perspective of the BMF model.
Figure 9 describes the extension of this new understanding to re-interpret
the simple case of tensile and compressive overloads and how their action may
be modeled. Consider the case of identical cycles AB, DE, and GH in Fig. 9(a)
and 9(b). We assume the stress ratio to be sufciently high in order to preclude
the possibility of crack closure. Near-tip cyclic stress strain response for these
cycles appears as Fig.9(c). We see that the baseline cycle AB would be associated
with a certain near-tip stress rA. If the stress ratio of this cycle had been higher
such that A and C were equal, the local stress would have risen to rC. In the
event of tensile overload as in Fig. 9(a), the following cycle ED sees a deep drop
in the near-tip stress to rE. However, if a compressive overload followed the tensile overload as in Fig. 9(b), the following cycle GH will see an increased local
mean stress due to the preceding yield in compression at F. The new mean
stress will still be lower than in AB. Assuming a unique relationship between the
near-tip mean stress and threshold stress intensity, the three identical load
cycles in question will follow different near-threshold da/dN curves as, indicated
in Fig. 9(d). Empirical determination of these modied da/dN curves of relevance to variable amplitude fatigue requires specially designed experiments
involving the controlled variation of near-tip mean stress. Note that, given the
impact of compressive overloads leading to increased tensile near-tip stresses,
there is no reason why the left extreme of these curves cannot tend towards
zero. Note also, that the right extreme for the da/dN curve is a high vacuum
response that effectively limits the extent to which compressive residual stress
can retard the fatigue process.
Figure 9(e) and 9(f) assist in understanding transients associated with neartip residual stress response after an overload. Here, A and C are the monotonic
baseline and overload plastic zones, respectively; D and B are the associated
cyclic plastic zones, respectively. Figure 9(f) shows the near-tip response at the
boundary of zone D. Since this point will see a fully elastic response, one may
assume that as the crack tip approaches this point, any hysteretic effects seen in
Fig. 9(c) will disappear. This means that beyond this point, from a residual
stress perspective, crack growth will be identical for the two cases in Fig. 9(a)
and 9(b). The memory about the compressive overload stands is erased from
this point. It also appears possible that, after some crack extension and well
before the boundary of C, near-tip stresses will be restored to the levels associated with the baseline conditions and one should, therefore, not see the extent
of the retardation zone assumed by the Wheeler and Willenborg models.
In summary, quite independent of crack closure, the residual stress effect is
manifested through the response of near-tip elements within the cyclic plastic
zone to stress history. Their response determines instantaneous DKth, suggesting that the near-threshold da/dN versus DK curve is not a material constant
and needs denition on a cycle-by-cycle basis as a function of near-tip mean or
maximum stress. As a consequence, the hysteretic near-tip stress variation induces cycle-sequence sensitivity in near-threshold crack extension, provided the
crack is fully open during the given cycle. In the event of partial crack closure,
cycle-sequence sensitivity is not possible because the minimum crack-tip stress

SUNDER, doi:10.1520/JAI103940 47

FIG. 9The new perspective of how tensile and compressive overloads distort the fatigue process. (a) Tensile overload; and (b) compressive overload following a tensile
overload. (c) Crack tip stress-strain response showing the effect of overloads on local
mean stress (crack-tip residual stress). Tensile overload pushes local stress into compression (ED), but if a compressive overload follows, local stress will rise (GH), though
not to the baseline value (AB). (d) Near threshold crack growth rates can swing dramatically depending on crack tip stress. (e) Overload cyclic plastic zone is small by comparison to the tensile overload plastic zone. Therefore, any sequence sensitive hysteretic
effect will disappear on its boundary, as seen in (f). This implies that beyond this point,
it will not matter whether a compressive overload followed the tensile one. However,
due to the combined action of closure and residual stress, most of the load-interaction
effect, bordering on crack retardation and possible momentary arrest, would have been
exhausted within the cyclic plastic zone. Conventional modeling techniques cannot
reproduce these effects because they ignore the cyclic plastic zone response and its effect
on threshold.

48 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

in the cycle is practically tied to the lowest possible crack-tip stress (see Fig.
8(e)).
Upon application of a tensile overload, the impact of the associated residual
compressive stress is immediate. This combines with the delayed development
of closure awaiting wake build up. As a consequence, retardation will be immediate in the event of the near-threshold response and delayed in the event of the
Paris Regime response or in high vacuum.34 Closure related retardation due to
overload will vanish only after the crack tip has extended well outside the overload plastic zone of the crack tip (see Fig. 6(d)). In contrast, the hysteretic nature of residual stress effects will disappear at the boundary of the overload
cyclic plastic zone and the retarding effect of residual stress will altogether disappear well before the crack tip exits the overload plastic zone as the near tip
stresses approach baseline values. This point has no connection with the point
where crack closure reaches its maximum. Thus, the combined action of crack
tip residual stress and closure will be limited in the crack extension interval.
However, over this small interval, retardation is likely to border on crack arrest.
The closure model accounts for only part of what happens except in the partial
case of Paris Regime growth rates. Additionally, the Willenborg and Wheeler
models altogether ignore the cyclic plastic zone response and treat the transient
process as a continuously changing one over the entire monotonic plastic zone.
The ramications of the deviation from reality of all existing approaches to
crack growth estimates under variable amplitude loading can be judged from
two important practical considerations of computation. First, the baseline
cyclic plastic zone where hysteretic effects dominate will be well under 10% of
the overload monotonic plastic zone size.35 Second, computed residual fatigue
life, being an integral of the growth rate function, will accumulate errors in
computed transient growth rates. This suggests the questionability of obtaining
reasonable crack growth estimates using available models. The suggestion may
appear preposterous when viewed against the operating framework of techniques currently in use to handle variable amplitude fatigue. An examination of
the empirical evidence and denition of the emerging perspective is, therefore,
pertinent.
The Experimental Evidence
A series of experiments were performed to verify each of the conclusions that
follow from interpreting variable-amplitude fatigue using the BMF model combined with closure. To avoid speculation, each experiment was specially
designed to deliver irrefutable fractographic evidence [5463]. These highlight

34)

Published fractographic data showing delayed retardation are restricted to the Paris
Regimethey show striations.
35)
Plastic zone size ratio is given as the square of the ratio of overload stress intensity to
half the baseline effective stress intensity range because cyclic plastic zone size is determined by twice the yield stress required for reverse yield.

SUNDER, doi:10.1520/JAI103940 49

quantiable effects for which there appears to be no alternate interpretation.36


The experiments imply the important criterion of falsiability.
It was a chance discovery that initiated this research in the late 1990s. Early
experiments on Al-alloy specimens using closure-free high stress ratio cycling
were performed in search of a correlation between the applied stress and short
crack response [54] (the so-called short crack effect). The experiments were performed under programmed loading with three steps of identical amplitude but
varying mean stress. Figure 4(b) shows a typical fatigue fracture from these
experiments. Clearly visible are sets of three bands of crack extension associated
with each set of three steps of loading. Note that each band is caused by a few
thousand load cycles and is not to be confused with striations from individual
cycles seen in the Paris Regime.
A noticeable difference in the crack growth rate is observed between individual steps at small crack size in Fig. 4(b). This difference gradually tapers out
to uniform crack growth rate as the crack grows much larger. The embedded
cycles were placed on the rising half of the major cycle.37 The authors correlated
measured crack growth rates with the maximum local notch root stress in individual steps and came up with an empirical equation to account for the short
crack effect as a function of local maximum stress [54]. At the time, this
approach was considered consistent with the prevailing notions of the so-called
short crack effect, where parameters such as local stress were considered
essential to explain what the stress intensity range could not. We did not consider the possibility that the steps with lower mean stress may experience the
benecial effect of preceding stressing at a higher level. We believed that having
ensured the crack was fully open by keeping stress ratios high, no load interaction effects were possible. Sometime after the publication of this work, a
chance38 discovery was made of equally spaced concentric circular bands
within voids on the fatigue fracture surface left behind by secondary particulates [55].
Several conclusions crucial to unraveling the nature of metal fatigue
emerged from the detailed study of fatigue voids. While it has long been known
that fatigue cracks form at the notch root, the new evidence conrms the possibility that early fatigue kinetics are the consequence of several competing mechanisms operating at different locations. At a high applied stress level promoting
intense reverse slip, the notch root surface is likely to develop several crack
36)

A few early experiments involved the analysis of striation patterns. The rest involved
estimates of spacing between marker bands employed to unambiguously characterize microscopic crack extension over thousands of near-threshold load cycles that cannot, in
their individual capacity, produce discernible growth marks. This technique permits
quantitative estimates of crack extension without a limitation on the minimum growth
rate. The pictures reproduced in this paper reach down to 108 mm/cycle.
37)
Had they been placed on the falling half, the retardation effect would have been much
more dramatic given the hysteretic response. At the time, the authors were not aware of
the phenomenon involved.
38)
In routine electron microscopy particulate voids are usually ignored as dark, featureless cavities.

50 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

origins almost simultaneously [61]. Plane stress conditions at the surface combined with assistance from the environment39 appear to dominate. With a
decreasing stress level, the number of such sites will diminish, with a general
tendency towards eventual sub-surface initiation.40 One may speculate that constraint in the interior will promote local defect growth by microscopic failure
through modes other than planar slip, which prefers plane stress conditions.
In Al-alloys, innumerable secondary particulates lying beneath the notch
root appear to bear evidence to the consequences of cyclic hydrostatic stresses
operating in the constrained region beneath the notch. These induce the gradual separation by interfacial fatigue cracking of the secondary particulate from
the matrix. Cyclic hydrostatic loading action is apparent from the simultaneous
onset and identical growth rate of typically six (even more in the case of the
irregular shape of the particulate) penny shaped interface cracks covering all six
sides of the particulate (see Fig. 10). The smallest crack size seen is of the order
of 0.125 lm, which may represent the smallest reproducible and traceable fatigue crack observed in research practice. The bands also indicate an incredibly
low growth rate down to 108 mm/cycle. The generally uniform spacing of the
concentric bands is of practical signicance, suggesting that the interfacial
crack growth rate appeared to be insensitive to change in the mean stress in
individual steps of the programmed load sequence employed. This was in contrast to the major short crack at the same proximity to the notch root! Surely,
the effect that caused growth rates to be different between steps in the major
short crack as seen in Fig. 4(b) ought to have also have inuenced the interfacial
crack growth! However, they apparently did not, after all.
There was, however, an important difference between the conditions under
which the two cracks grew. Unlike the major crack originating from the surface
and continuously exposed to the environment, interfacial cracks around secondary particulates grow in ideal vacuum. This is conrmed by simultaneous
cracking around the particulate that could not have progressed without cyclic
hydrostatic tensile stresses, and these in turn will disappear once the particulate
is exposed and constraint disappears. There was obviously something linked
not with the macro-mechanics of the notch response, but rather, with the
micro-mechanism of crack extension that seemed to determine fatigue resistance. A possibility has now emerged that vacuum inhibits the root cause for the
mean (residual) stress effect in metal fatigue. It was also possible that in air, it
was not the applied mean stress itself, but the sequence of its change (load history) that was responsible. Perhaps, indeed, vacuum does disable residual
(mean) stress related effects?
Reference [56] describes an experiment dedicated to conclusively isolate
the role of environment in near-threshold fatigue by falsication. The experiment involved testing to failure under the same three-step programmed loading,
but alternating between air and vacuum every given number of blocks. The

39)
In Al-alloys, interfacial environmental attack causes early pitting through the separation of secondary particulates on the notch surface. Each pit is a potential initial defect.
40)
Gigacycle fatigue is almost always associated with internal crack formation.

SUNDER, doi:10.1520/JAI103940 51

FIG. 10Fatigue voids and microvoids [55]. (a) Proof that individual voids seen on Alalloy fatigue fractures were formed by fatigue-separation of secondary particulates from
the matrix and not due to high Kmax quasi-static failure as claimed in [40]. Evidence of
interfacial cracking under three-step programmed loading (inset). Clear, equally spaced
bands marked by marker loads between steps indicate that the change in the mean
stress level did not have any effect on the crack extension due to the 2000 cycles in each
step. The schematic shows cyclic hydrostatic forces responsible for the cracking. (b)
Rare picture of the secondary particulate that remained bonded to the fatigue fracture.
The area immediately around the particulate is evidently formed by fatigue. The surrounding area is marked by clusters of microvoids that coalesced to cause quasi-static
crack extension. Microvoids are formed by very high hydrostatic stresses leading to
microcavitation, with the walls between cavities failing in ductile fashion due to localized plane stress conditions. Note the vast difference in size between particulate voids
and microvoids, indicating that one cannot be confused with the other (as was the case
in [40]). (c) Multiple interfacial cracks separating an irregularly shaped particulate sitting on the boundary of three grains suggesting the action of tensile cyclic hydrostatic
stress.

52 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

vacuum steps were twice the cycle count to account for retarded growth. The
switch from vacuum to air took a mere few minutes. However, the switch from
air to high vacuum (108 Torr) required more than 48 h, with the entire experiment lasting several weeks. The experiments provided conclusive evidence in
support of the BMF model (see Fig. 11(a) and 11(b)). In air (Fig. 11(a), left), the
notch root small crack growth rate in the three steps varied exactly as in the earlier experiment (Fig. 4(b)). However in high vacuum, the three steps caused
identical crack extension as seen at top right of Fig. 11(a) and magnied as in
Fig. 11(b). The instant air was released into the chamber, and the growth rates
in the three steps once again became different. This conrmed the absence of
the crack-tip residual stress effect in high vacuum. It also provides an alternate
explanation for the so-called short crack effect.
If, indeed, residual stress operates by the moderation of cycle-by-cycle environmental action, it should reproduce on all metals and in the presence of any
active species that can diffuse and thereby adversely affect fatigue resistance.
To conrm this possibility, experiments were repeated on a Ni-base superalloy
at an elevated temperature, once again on the same machine, in air and in vacuum. In this case, the results were even more dramatic, apparently because of
the sensitivity of the crack tip oxidation to the near-tip residual stress [58].
If, indeed, the near-tip stresses within the cyclic plastic zone control diffusion
kinetics and through it, near-threshold crack extension, they should exhibit hysteretic sequence sensitivity. This hypothesis was successfully veried by tests performed under two different programmed sequences, one directed at growth rates
closer to the Paris Regime and another, closer to threshold [57]. Figure 11(c)
shows a typical fractograph obtained from the second experiment performed
using the sequence shown in the inset. The three steps are of identical small amplitude set way above expected closure levels, in order to induce hysteretic neartip stress variation between steps 1 and 3, as shown schematically in Fig. 8(d). As
expected, the growth rate in step 3 is dramatically retarded by comparison to step
1. If the same experiment were to be performed in high vacuum, the crack extension would be identical in all three steps and close to that in step 3.
Finally, another experiment was designed; this time, to demonstrate the
synergy of crack closure and the residual stress effect [58]. The results are
briey summarized in Fig. 12. The load sequence was specially designed to
selectively induce full crack closure, partial crack closure, and a fully open
crack. A key-hole notched C(T) specimen was chosen to induce natural crack
formation under conditions of notch root compressive residual stress due to
monotonic yield at maximum stress. Notch root crack closure is known to be
sensitive to local residual stress [59]. Steps of identical small amplitude were
embedded at three different mean stress levels on the rising and falling half of
the major cycle. The fractographs provide a graphic illustration of how the
notch root residual stress affects crack closure and how crack closure combines
with crack-tip residual stress effects to control variable amplitude fatigue crack
growth. Initial notch root yield in tension induced residual compressive stress
that reduced local stress ratio and thereby increased closure levels in the initial
stage of fatigue when the crack was barely 0.05 mm deep. As a consequence,
steps 1and 5 were fully closed and steps 2 and 4 partially closed. Furthermore,

SUNDER, doi:10.1520/JAI103940 53

FIG. 11(a) and (b) Proof of residual stress operative mechanism by falsication. (a)
Crack growth under three step programmed loading in air and high vacuum (top right)
[56]. Noticeable retardation in crack extension in the second and third step is reproduced across multiple blocks. However, upon switching to high vacuum (top right and
magnied picture (b), the crack extension in all three steps is identical. The switch in
growth rate response was instantaneous in both the air-vacuum and vacuum-air transitions suggesting the virtual absence of any transient effects and also the impossibility
of crack closure playing a role. The Wheeler, Willenborg, and closure models cannot
explain these observations. (c) Proof of the effect of the hysteretic crack-tip stress-strain
response on atmospheric near-threshold crack growth rate. Note the substantial retardation in step 3 because of compressive crack-tip stresses due to load cycles lying on the
falling half of the major cycle, as explained in Figs. 8(c), 9(d). This effect tapers out into
the Paris Regime, a phenomenon that the Wheeler/Willenborg and closure models cannot simulate.

54 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

FIG. 12Experiment on Al-alloy to demonstrate the synergy of the transient notch root
crack closure and residual stress [58]. (a) Multi-step programmed load sequence
designed to induce hysteretic residual stress variation in steps 2 and 4. Note that the duration of step 3 is half of the others. Selected max load induced notch root tensile yield
leaving compressive residual stress at the notch root. (b) Macro showing the notch root
at left and the locations of fractographs c and d. (c) Identical growth from steps 2 and 4
indicates partial crack closure (at about 50% stress) and also explains why steps 1 and
5 did not extend the crack. (d). Almost identical growth in 2 and 4 and equal growth in
1 and 5 suggests that the closure level was around 40%. (e) A large difference in the
crack extension between 2 and 4 suggests a noticeable hysteretic variation in crack tip
mean stress. The closure level must have dropped to the long crack level of 30% (crack
size 1.5 mm). However, steps 1 and 5 are partially closed, causing equal crack extension. No model or software in commercial use today is capable of simulating the crack
extension patterns shown.

SUNDER, doi:10.1520/JAI103940 55

in Fig.12(c), we see equal bands from 2,4 and no crack extension during 1 and 5.
Figure 12(d) shows a fractograph at a location about 0.2 mm from the notch
root, where closure level has by now dropped somewhat. As a consequence,
steps 1 and 5 are partially open and steps 2 and 4 are fully open. This is indicated by discernible and equal crack extension in steps 1 and 5 and marginal retardation in step 4 by comparison to 2 due to the hysteretic difference in neartip stress. Figure 12(e) shows a fractograph from a location about 1.5 mm from
the notch root where closure has dropped to a long crack level of about 30%.
This causes equal crack extension in steps 1 and 5 and considerable retardation
in step 4 by comparison to step 2. Also note that the crack extension in step 2 is
retarded by comparison to step 3 (which is of half the duration). If this test were
to be conducted in high vacuum, crack extension in steps 2 and 4 would have
been identical and exactly twice that in step 3 (merely because of twice the cycle
count). Crack growth in steps 1 and 5 would have remained less, due to partial
crack closure.
Implications of Empirical Evidence
Some 150 years after the effect was rst noticed, the underlying science behind
variable-amplitude fatigue is nally unraveling. The connection between cracktip hydrostatic stress and near-threshold growth rates resulting in DKth ceasing
to be a material constant completely changes ones perspective of laboratory
test data and their relevance to engineering applications. In view of its sensitivity to hysteretic crack-tip stress-strain response, DKth emerges, in effect, as the
sole reason for fatigue being cycle-sequence sensitive. For this reason, DKeff can
no longer be treated as a similarity criterion for near-threshold fatigue crack
growth. Crack closure can no longer be assumed to account by itself for stress
ratio effects and for load interaction phenomena. Retardation is possible without closure. A fully open crack can accelerate under conditions of increased tensile residual stresses at the crack tip. As it turns out, all these are possible
merely because of the moisture in the air.
Interestingly, Marci and Lang came close to intuitively judging the signicance of crack-tip hysteretic stress-strain [64]. They proposed the idea of a minimum K called KPR (for propagation) that needs to be exceeded by DKth, for the
onset of fatigue crack extension. A simple yet time-consuming experimental
technique was developed to determine KPR, by cycling at DKth (treated as a material constant) with gradually stepped up Kmin until the detectable onset of
crack extension. In simple variable-amplitude experiments on Al-alloys that
were later extended to other alloys, it was found that KPR was sensitive to load
history and as a rule, exceeded Kop. It was also found that depending on load
history, KPR changed in ways that could not be explained by closure. On the basis of these two considerations and on the basis of KPR capability for improved
estimates of spectrum load crack growth, the authors concluded that in KPR, a
variable had been found that is an effective replacement for Kop. They specically noted its ability to not only account for closure, but also for a certain
notional residual stress effect at the crack tip. Notional, because its potential for
cycle-by-cycle hysteretic variation and the signicance of the cyclic plastic zone

56 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

were both ignored. Unfortunately, Marci and Lang failed to realize that KPR
may have been actually accounting for the instantaneous change in DKth, which
they had wrongly assumed to be a material constant. The difference KPR Kop,
may, in fact, represent instantaneous DKth changing as a function of load history. One may speculate that the authors would have found this value to remain
virtually constant in high vacuum and equal to the difference in DKth between
vacuum and air! Like many others before them, Lang and Marci appear to have
succumbed to the perception of DKth as a material constant.
This new understanding nally allows for reassessment of the residual
stress effect by separating its mechanics from operating mechanisms. Residual
stress in the cyclic plastic zone ahead of the crack tip is controlled primarily by
load history and the associated cycle-by-cycle stress intensity sequence. By virtue of its immediate proximity, crack tip surface resistance to fracture is directly
affected in atmospheric near-threshold fatigue. On the contrary, the notch root
residual stress, and other such (remote) macroscopic stress distributions control local stress ratio and through it, crack closure [59, 65]. Previous work on
the subject may have offered powerful tools to address the mechanics of fatigue
and fracture mechanics, yet they did not have the benet of clarity in scientic
understanding, without which realistic analytical modeling or even targeted experimental research appears rudderless.
Emerging Avenues for Future Work
Historically, research on the residual stress effect was restricted to technologies
for its introduction, for its removal, and for its measurement. Its operating
mechanism in moderating fatigue damage seems to have elicited little intellectual curiosity. The discovery of the science behind how near-tip residual stress
affects near threshold fatigue crack growth and its distinction from how the
notch root residual stress affects crack closure together open avenues for future
research towards improved modeling in engineering applications. Listed in the
following sections are a few such avenues.
Crack closureOf the different experimental techniques available, only
fractography [37] and crack tip laser interferometry [36] appear to deliver
authentic measurements of crack closure. Of these, the rst requires special
load sequences and cumbersome microscopy and is restricted to materials that
are fractography friendly. Furthermore, since crack extension is also sensitive
to crack-tip residual stress, one must ensure in designing the load sequence,
that the crack-tip stress remains unchanged at the applied maximum stress.
Failure to ensure this condition can lead to distortions as seen in [58]. The second technique is expensive and demands special equipment and in the end, only
provides surface measurements. All other techniques deliver far eld measurements that cannot possibly serve as a reliable measure of a near-eld phenomenon. Until laboratory techniques become available that can plot local strain
within or very close to the cyclic plastic zone versus the applied load, it is
unlikely that one can make useful measurements of crack closure in routine

SUNDER, doi:10.1520/JAI103940 57

testing practice. The development of such techniques appears to hold much


promise in future applications.
Elaborate nite-element solutions have been developed to compute stress
intensity for a given crack path, given applied loading, and a given resident residual stress eld in a material including processing induced elds, elds due to
interference t fasteners, along with response induced elds such as at notches
subject to local yield. State-of-the-art numerical simulation as demonstrated by
Seshadri et al. [66,67] appears to provide realistic cycle-by-cycle estimates of
the stress-strain response at any point in a cracked structure, under any given
load sequence. This includes the cyclic inelastic response around the crack tip
in the presence of crack closure, which appears to conrm that closure is a
mechanics driven phenomenon involving crack wake development into the
monotonic plastic zone, applied cyclic loading, and possible residual stress
elds in the material. The crack-tip load-displacement data reported in Refs.
[66, 67] appear to accurately reproduce the laser interferometry measurements
shown in Fig. 6(e). They clearly show the hysteretic response and the associated
loop formation in displacement versus load. Closure load can be unambiguously associated with loop closure. Yet, the authors have preferred to interpret
closure in terms of change in compliance response as per ASTM E647, leading
to exaggerated closure estimates.
One may conclude that contradictory measurements and estimates of crack
closure stress may, in part, be attributed to the selection of compliance offset
points in measurements, that is synonymous with wake contact point in analyses. In such a denition, the focus is deected from the primary objective, which
is to dene which fraction of the applied load cycle is responsible for creating
the observed cyclic plastic zone size.
Threshold Stress IntensityThe association of threshold fatigue with the
BMF, crack tip diffusion, and reaction kinetics implies that the highest possible
DKth will be in high vacuum and may be treated as a material constant. This parameter characterizes the maximum benecial value that compressive residual
stresses can have on fatigue in air. The ultimate goal would be to come up with
an analytical model relating DKth to crack-tip diffusion kinetics as a function of
history-sensitive instantaneous hydrostatic stress ahead of the crack tip.
An intermediate objective may be to characterize DKth under a variety of
controlled near-tip stress conditions, while at the same time ensuring the absence of closure. If closure is present, an error in its measurement will carry
over to the DKth estimate. Such errors are unacceptable in threshold studies
because they may be of the same order as DKth. By keeping the stress ratio sufciently high, closure free DKth measurements are possible.41 An exploratory
study of Hi-R DKth under the action of periodic overloads provided a linear relationship between the overload plastic zone ratio and closure-free DKth [61]. The
41)
Actual mid-thickness closure levels seldom exceed 25-30% of the max load under constant amplitude loading, when measurements are made using techniques such as fractography or laser indentation interferometry.

58 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

study on two Al-alloys conrmed that variable-amplitude DKth does indeed


approach vacuum levels. It is proposed to repeat the experiments with compressive overloads in order to determine the lower bound of DKth at a given test frequency. Experiments are also proposed to examine the effect of the hysteretic
near-tip response on thresholds. Finally, hold times at periodic tensile and compressive overloads are also likely to affect the atmospheric threshold fatigue
response because they induce stress relaxation.
Cyclic Stress-Strain ResponseThe modeling effort requires answers to new
questions. Is there a characteristic distance behind the crack tip whose stressstrain response controls diffusion kinetics? What stress-strain curve would apply
at this point, considering constraint to be a controlling factor? Recent research
[68] reconrms diffusion and reactivity of active species such as moisture as the
source of accelerated near-threshold fatigue response. An inter-disciplinary
effort to connect the crack-tip cyclic mechanical response to diffusion and chemical reaction kinetics may be the next step towards modeling near-threshold fatigue resistance in variable-amplitude fatigue. Determining the connection
between DKth, near-tip constraint, and cyclic strain hardening coefcient, combined with the subsequent incorporation of sensitivity to hold time serve as
attractive long-term goals to tie in material cyclic stress-strain response with fatigue. It is likely that materials with low strain hardening properties and reduced
constraint will exhibit a reduced sensitivity of DKth,eff to the stress ratio. This is
because increasing the stress ratio may not induce much increase in the near-tip
stress. However, all materials will exhibit a hysteretic stress-strain response and
will therefore exhibit stress history effects. Obviously, experiments restricted to
constant amplitude loading are unlikely to carry much practical value, apart
from underscoring the signicance of the phenomenon as in [68].
The emergence of MEMS and biomedical applications of metallic components and the application of nano-structured materials holds much scope for
the application of future work because of the potential dominance of surface
phenomena in these cases.42 Finally, the unication of near-threshold variableamplitude fatigue at room and elevated temperatures through the near-tip
response holds promise in gas and steam turbine applications. It is likely to
assist in improved modeling of the effect of overloads at high temperature,
hold-time, and creep-fatigue interaction effects. For example, it now appears
obvious that crack extension during hold at a given load at an elevated temperature will be driven by diffusion kinetics moderated by crack-tip residual stress.
This opens up the possibility of modeling the interaction of overloads with the
hold-time.
The bulk of fatigue damage due to many spectra including transport aircraft load spectra is from the smallest load cycles that arguably advance fatigue
through near-threshold mechanisms yet are subject to the history effects from
periodic overloads. It is plausible that the B737 Aloha Airlines incident and,
42)

For a given volume, the total exposed surface area increases with the decreasing size
and scale of constituents.

SUNDER, doi:10.1520/JAI103940 59

more recently, with a Southwest Airlines fuselage panel may have been associated with stress intensity ranges deemed to be sub-threshold from laboratory
test data on coupons tested at a higher frequency. In the course of about fteen
years of service, such aircraft would experience over 80,000 ights or 107 small
load cycles.
Overall, modeling of the residual stress effect holds the promise of advancements in the quality of fatigue life estimates with a greater reliance on simulation and a reduced emphasis on expensive empirical inputs. Lack of it will
continue to force dependence either on corrections of cumulative damage to
match experimental data, or on corrections of the crack driving force to compensate for the inability to account for the change in material resistance. In the
meantime, disciplines other than fatigue and fracture mechanics will continue
to determine the safety and durability of engineered products, while in the long
term, metal fatigue may be simply rendered less relevant by advances in the
application of engineered composites that would be immune to the type of
mechanisms that induce metal fatigue.

Summary
1. The practical relevance of cyclic-slip to metallic component durability
is overrated. Slip-driven fatigue dominates low-cycle fatigue and crack
growth at rates exceeding 104 mm/cycle. In durable fatigue designs
most of the fatigue life is expended at crack growth rates below the
Paris Regime. Atmospheric metal fatigue under these conditions is
controlled by the near-threshold response, where the consequences of
cyclic crack-tip surface activity overshadow the possible consequences
of cyclic slip.
2. Crack-tip surface activity progresses during each rising load half-cycle
with rising near-tip stress acting as a diffusion pump to promote
embrittlement or chemical weakening of surface atomic layers and
associated accelerated crack extension. At ambient temperature, reaction with moisture releases hydrogen for diffusion. At elevated temperature, oxidation is involved. The depth and extent of such an attack is
moderated by local hydrostatic stress, that in turn, is determined by
the stress ratio and cycle-sequence sensitive near-tip residual stress.
The effect is restricted to crack-tip surface atomic layers and therefore
becomes insignicant as the growth rate progresses into the Paris Regime. It is totally absent in high vacuum.
3. For the purpose of understanding its effect on metal fatigue, residual
stress may be divided into remote (or macroscopic, or crack-free)
stress distribution and the local (microscopic) eld associated with the
crack tip response.
4. The macroscopic eld, including residual stresses left by mechanical
processing, and those induced by local inelastic static or cyclic
response such as at notches, control the local stress ratio and associated fatigue crack closure in conjunction with applied cyclic load

60 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

5.

6.

7.

8.

9.

10.

11.

conditions and crack wake development into the monotonic plastic


zone. These can be computed using state-of-the-art analytical tools
that determine the material and structural stress-strain response along
with the stress intensity function for a given crack size, shape, and
path.
Crack closure can be unambiguously determined only from the shape
and closure of the loop formed either by the inelastic near-tip strain,
or, by non-linear hysteretic near-tip wake displacement when plotted
against the applied load. The conventional approach of monitoring
wake contact therefore appears misdirected and may be the root cause
of incorrect closure estimates reported in the literature, be it computed
values, or those obtained experimentally using standard practices such
as ASTM E-647.
The microscopic crack-tip stress eld is the result of the action of the
next loading cycle, superposed on the residual crack-tip stress-strain
eld at the end of the previous unloading half cycle. This eld will
reect the effect of the macroscopic residual stress distribution in the
material (through stress intensity), along with that of the monotonic
plastic zone.
In atmospheric fatigue, by shifting the near-tip stress up or down, the
microscopic eld moderates diffusion kinetics to determine instantaneous threshold stress intensity, strictly speaking, for a given ambient
partial pressure of active species (humidity), temperature, and cycling
frequency. Vacuum threshold stress intensity serves as its upper limit.
There is no known lower limit.
The microscopic eld will exhibit a signicant cyclic inelastic hysteretic response while the crack is fully open. The associated change in
the threshold stress intensity is the root cause for cycle-by-cycle (hysteretic) load sequence sensitivity in variable-amplitude metal fatigue.
This component of sequence sensitivity will be absent in high vacuum
and diminish in air into the Paris Regime because its effect is restricted
to crack-tip surface atomic layers.
Conventional modeling of notch fatigue under variable amplitude loading using the local stress-strain (LSS) approach does not carry any scientic rationale. If it does correctly describe trends in fatigue response,
it is by the coincidental qualitative similarity of the cyclic notch root
and crack-tip response. This similarity vanishes under a fully elastic
notch root response, exposing the invalidity of the LSS approach.
Commercially available models of variable amplitude fatigue crack
growth, including the Wheeler and Willenborg models, focus on the
monotonic plastic zone and therefore, essentially address crack closure. These models treat threshold stress intensity as a material constant even under variable amplitude loading. Therefore, correct
estimations by such models of the atmospheric variable amplitude fatigue growth rate below 104 mm/cycle can only be by accident.
For a given applied load cycle, crack front orientation, and tortuosity
moderate crack-tip stress-strain response, crack closure determines its

SUNDER, doi:10.1520/JAI103940 61

effective magnitude and the near tip stress response superposed on residual stress determines instantaneous resistance (threshold stress intensity). The rst is sensitive to the crack extension history. The second
is sensitive to the crack extension and loading history. The third is sensitive to the load cycle-sequence and loading history. Variable-amplitude
fatigue response needs to be modeled as the synergy of all three.
12. Further improvements to analytical modeling of variable-amplitude
fatigue demand consideration of threshold stress intensity as a cyclesequence sensitive variable. They would also benet from reliable laboratory measurements and analytical estimates of crack closure and
from improved characterization of the crack-tip response to variations
in crack front geometry (shielding effects). Such studies should include
the effect of constraint.

Acknowledgments
Some of the experiments and all of the reported fractography were performed
at the Air Force Research Laboratories (AFRL), WPAFB, OH, USA. Other
experiments were performed at BiSS Research, Bangalore. The author deeply
appreciates the support and encouragement provided by colleagues in both laboratories and also the University of Dayton Research Institute (UDRI).
References
Thorneycroft, T., On the Form of Shafts and Axles, Proceedings of the Institution
of Mechanical Engineers, London, Oct 1850, pp. 3541 pp. 415.
[2]
Braithwaite, F., On the Fatigue and Subsequent Fracture of Metals, Proc. Inst.
Civ. Eng, London, May 1854.
[3]
Wohler, A., Uber die Festigkeitsversuche mit Eisen und Stahl, Berlin, Ernst und
Korn, 1870.
[4]
Anon, Wohlers Experiments on the strength of Metals, Engineering, Vol. 4, 1867,
pp. 160161.
[5]
Manson, S. S., Future Directions for Low Cycle Fatigue, Low Cycle Fatigue, ASTM
Spec. Tech. Publ. 942, H. D. Solomon, G. R. Halford, and B. N. Leis, Eds., American
Society for Testing Materials, Philadelphia, 1988, pp. 1539.
[6]
Miner, M. A., Cumulative Damage in Fatigue, Trans. ASME J. Appl. Mech., Vol.
12, 1945, pp. A159A164.
[7]
Anon., Wohlers Experiments on the Fatigueof Metals, Engineering, June 1871,
pp. 199441.
[8]
Bauschinger, J., On the Change of the Elastic Limit and Strength of Iron and Steel
by Tension and Compression, by Heating and Cooling and by Often Repeated
Loading, Technical Report, Munich Technical Univ., Munich, Germany, 1886 (in
German).
[9]
Goodman, J., Mechanics Applied to Engineering, Longmans-Green, London, 1899.
[10] Raju, K. N., Workshop on Fatigue, Fracture and Failure Analysis, Notes, Vol. 1,
National Aeronautical Laboratory, Bangalore, March 1979.
[11] Hull, D., Bacon, D. J., Introduction to Dislocations, Fourth Edition, ButterworthHeinemann, Oxford, 2001.

[1]

62 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

[12]

[13]
[14]
[15]
[16]

[17]
[18]
[19]
[20]

[21]

[22]

[23]

[24]
[25]

[26]

[27]
[28]
[29]
[30]
[31]

[32]

Grosskreutz, J. C., A Critical Review of Micromechanisms in Fatigue, Fatigue


An Interdisciplinary Approach, Syracuse University Press, Syracuse, 1964, pp. 27
62.
Schijve, J., Fatigue of Materials and Structures, Springer, New York, 2003, p. 317.
Gassner, E., Strength Experiments Under Cyclic Loading in Aircraft Structures,
Luftwissen Vol. 6, 1939, pp. 6164 (in German).
Neuber, H., 1937, Kerbspanmungslehre [Theory of notch stresses], J.W. Edwards,
Ed., Springer, Berlin, Ann Arbor, MI, 1946.
Ramberg, W. and Osgood, W. R., Description of Stress-Strain Curves by Three
Parameters, Technical Report No. 902, National Advisory Committee for Aeronautics, Washington, D.C., 1943.
Wetzel, R. M., 1971, A Method for Fatigue Damage Analysis, Ph.D. thesis, Dept.
of Civil Engineering, Univ. of Waterloo, ON, Canada.
SAE Fatigue Design Handbook, 3rd ed., AE-22, Society for Automotive Engineers,
Warrendale, PA, 1997).
Matsuishi, M. and Endo, T., Fatigue of Metals Subjected to Varying Stress, Trans.
Jpn. Soc. Mech. Eng., March, 1968.
van Dijk, G. M. and de Jonge, J. B., Introduction to a Fighter Aircraft Loading
Standard for Fatigue Evaluation FALSTAFF, Report No. NLR MP 75017U,
National Aerospace Laboratory, Amsterdam, The Netherlands, 1975.
de Jonge, J. B., Schutz, D., Lowak, H., and Schijve, J., A Standardized Load
Sequence for Flight Simulation Tests on Transport Aircraft Wing Structures,
Report No. NLR TR 73029U, National Aerospace Laboratory (NLR), Amsterdam,
The Netherlands, 1973.
Sunder, R., Rainow Applications in Accelerated Cumulative Fatigue Analysis,
The Rainow Method in Fatigue, Y. Murakami, Ed., Butterworth -Heinemann,
Oxford, 1992, pp. 6776.
Neggard, G. R., The History of the Aircraft Structural Integrity Program, Report
No. 680.1B, Aerospace Structural Information and Analysis Center (ASIAC), Flight
Dynamics Directorate, WPAFB, Ohio 1980.
Schijve, J., Cumulative Damage Problems in Aircraft Structures and Materials,
Aeronaut. J., Vol. 74, 1970, pp. 517532.
Sunder, R., Porter, J., and Ashbaugh, N. E., The Effect of Stress Ratio on Fatigue
Crack Growth Rate in the Absence of Crack Closure, Int. J. Fatigue, Vol. 19, 1997,
pp. S211S221.
Sunder, R., Contribution of Individual Load Cycles to Fatigue Crack Growth
under Aircraft Spectrum Loading, M. R. Mitchell and R. W. Landgraf, ASTM Spec.
Tech. Publ., Vol. 1122, 1992, pp.176190.
Paris, P. C., Gomez, M. P., and Anderson, W. E., A Rational Analytical Theory of
Fatigue, Trend Eng., Vol. 13, 1961, pp. 914.
Figge, I. E., and Newman, J. C., Fatigue Crack Propagation in Structures with
Simulated Rivet Forces, ASTM Spec. Tech. Publ., Vol. 415, 1967, pp. 7193.
Schijve, J., Effect of Load Sequences on Crack Propagation Under Random and
Program Loading, Eng. Fracture Mech., Vol. 5, 1973, pp. 269280.
Wheeler, O. E., Spectrum Loading and Crack Growth, ASME J Basic Eng., Vol.
94, 1972, pp. 181186.
Willenborg, J., Engle, R. H., and Wood, H. A., A Crack Growth Retardation Model
Based on Effective Stress Concepts, Report No. AFFDL-TM-71-1 FBR, WPAFB,
OH, 1971.
Walker, E. K., Effects of Environment and Complex Load History on Fatigue
Life, ASTM Spec. Tech. Publ., Vol. 462, 1970, pp. 114.

SUNDER, doi:10.1520/JAI103940 63

[33]
[34]

[35]

[36]

[37]
[38]
[39]
[40]

[41]
[42]
[43]
[44]
[45]
[46]
[47]

[48]

[49]

[50]

[51]

Elber, W., The Signicance of Fatigue Crack Closure, ASTM Spec. Tech. Publ.,
Vol. 486, 1971, pp. 230242.
Newman, J. C., A Crack-Closure Model for Predicting Fatigue Crack Growth Under
Aircraft Spectrum Loading. Methods and Models for Predicting Fatigue Crack Growth
Under Random Loading, ASTM Spec. Tech. Publ., J.B. Chang and C. M. Hudson,
Eds., Vol. 748, 1981, pp. 5384.
de Koning, A. U., A Simple Crack Closure Model for Prediction of Fatigue Crack
Growth Rates Under Variable-Amplitude Loading, Fracture Mechanics, ASTM
Spec. Tech. Publ., R. Roberts, Ed., Vol. 743, 1981, pp. 6385.
Ashbaugh, N. E., Dattaguru, B., Khobaib, M., Nicholas, T., Prakash, R. V., Ramamurthy, T. S., and Seshadri, B. R., Experimental and Analytical Estimates of Fatigue Crack Closure in an AluminumCopper Alloy. Part I. Laser Interferometry
and electron Fractography, Fatigue Fract. Eng. Mater. Struct., Vol. 20(7), 1997, pp.
951961.
Sunder, R., and Dash, P. K., Measurement of Fatigue Crack Closure Through Electron Microscopy, Int. J. Fatigue, Vol. 4, April 1982, pp. 97105.
Schijve, J., Four Lectures on Fatigue Crack Growth, Eng. Fracture Mech., Vol. 11,
1979, pp. 176221.
Sunder, R., A Unied Model of Fatigue Kinetics Based on Crack Driving Force and
Material Resistance, Int. J. Fatigue, Vol. 29, 2007, pp. 16811696.
Riddell, W. T. and Piascik, R. S., Stress Ratio Effects on Crack Opening Loads and
Crack Growth Rates in Aluminum Alloy 2024, Fatigue Fracture Mechanics, ASTM
Spec. Tech. Publ., Vol. 1332, T. L. Panontin and S. D. Sheppard, Eds., Vol. 29, American Society for Testing and Materials, West Conshohocken, PA, 1999, pp. 40725.
Forsyth, P. J. E., Fatigue Damage and Crack Growth in Aluminium Alloys, Acta
Metall., Vol. 11, 1963, pp. 703719.
Sunder, R., Binary Coded Event Registration on Fatigue Fracture Surfaces, J.
Soc. Env. Engrs., SEECO, London, 1983, p. 197.
Laird, C., Mechanisms and Theories of Fatigue, Fatigue and Microstructure, ASM,
Metals Park, OH,1978, pp. 149204.
Zhang,. J. Z., A Shear Band Decohesion Model for Small Fatigue Crack Growth in
an Ultra-Fine Grain Aluminum Alloy, EFM, Vol. 65, 2001, pp. 665681.
Sunder, R., Fatigue as a Process of Brittle Micro-Fracture, FFEMS, Vol. 28(3),
2005, pp. 289300.
Gsell, M., Jakob, P., and Menzel, D., Effect of Substrate Strain on Adsorption, Science Vol. 280, 1998, pp. 717720.
Ro, Y., Agnew, S. R., and Gangloff, R. P., Environmental Fatigue-Crack Surface
Crystallography for Al-Zn-Cu-Mg-Mn/Zr, Metall. Mater. Trans. A, Vol. 39A, 2008,
pp. 14491465.
Schijve, J. and Arkema, W. J., Crack Closure And the Environmental Effect on
Fracture Mode Transition in Fatigue Crack Growth, Report No. VTH-217, Delft
Univ., Delft, The Netherlands, 1976.
Vogelesang, L. B. and Schijve, J., Environmental Effects on Fatigue Failure Mode
Transition Observed in Aluminium Alloys Report No. LR-289, Delft Univ. of Technology, Delft, The Netherlands, 1979.
Gach, E. and Pippan, R., Cyclic Crack Tip Deformation the Inuence of Environment, Proceedings of the Tenth International Conference on Fracture, International
Congress of Fracture, Hawaii, Dec 2001 [Paper ICF 100420OR].
Bowles, C. Q., 1978, The Role of Environment, Frequency and Wave Shape During
Fatigue Crack Growth of Aluminum Alloys, Ph.D. thesis, Report No. LR-270, Delft
Univ. of Technology, Delft, The Netherlands.

64 JAI  STP 1546 ON FATIGUE AND FRACTURE MECHANICS

[52]

[53]

[54]

[55]
[56]
[57]
[58]

[59]
[60]

[61]

[62]

[63]

[64]
[65]

[66]

[67]

[68]

Petit, J., Henaff, G., and Sarrazin-Baudoux, C., Mechanisms and Modeling of
Near-Threshold Fatigue Crack Propagation, Fatigue Crack Growth Thresholds, Endurance Limits and Design, ASTM Spec. Tech. Publ., J. C. Newman, Jr. and R. S.
Piascik, Eds., Vol. 1372, American Society for Testing and Materials, West Conshohocken, PA, 2000.
Bradshaw, F. J. and Wheeler, C., The Effect of Gaseous Environment and Fatigue
Frequency on the Growth of Fatigue Cracks in Some Aluminium Alloys, Int. J
Fract. Mech., Vol. 6, 1969, pp. 255268.
Sunder, R., Porter, W. J., and Ashbaugh, N. E., Stress-Level Dependent Stress Ratio
Effect on Fatigue Crack Growth, Fatigue and Fracture Mechanics: Twenty-Ninth
Volume, ASTM Spec. Tech. Publ., Vol. 1332, T. L. Panontin and S. D. Sheppard,
Eds., American Society for Testing and Materials, West Conshohocken, PA, 1999.
Sunder, R., Porter, W. J., and Ashbaugh, N. E., Fatigue Voids and Their Signicance, Fatigue Fract. Eng. Mater. Struct., Vol. 25, 2002, pp. 10151024.
Sunder, R., Porter, W. J., and Ashbaugh, N. E., The Role of Air in Fatigue Load
Interaction, Fatigue Fract. Eng. Mater. Struct., Vol. 26, 2003, pp. 116.
Sunder, R., On the Hysteretic Nature of Variable-Amplitude Fatigue Crack
Growth, Int. J. Fatigue, Vol. 27, 2005, pp. 14941498.
Sunder, R., Fractographic Reassessment of the Signicance of Fatigue Crack Closure, Fatigue and Fracture Mechanics, ASTM Spec. Tech. Publ., Vol. 1461, S. R.
Daniewicz, J. C. Newman, and K. H. Schwalbe, Eds., American Society for Testing
Materials, Philadelphia, Vol. 34, 2005, pp. 2239.
Anandan, K. and Sunder, R., Closure of Part-Through Cracks at the Notch Root,
Int. J. Fatigue, Vol. 9, 1987, pp. 217222.
Ashbaugh, N. E., Porter, W. J., Rosenberger, A. H., and Sunder, R., EnvironmentRelated Load History Effects in Elevated Temperature Fatigue of a Nickel-Base
Super-Alloy, Proceedings Fatigue, Stockholm, Sweden, June 2-7, 2002, EMAS (2002).
Sunder, R., Effect of Periodic Overloads on Threshold Fatigue Crack Growth in
Al-Alloys, Fatigue Fracture Mechanics, ASTM Spec. Tech. Publ., S. R. Daniewicz, J.
C. Newman, K. H. Schwalbe, Eds., Vol. 1461, American Society for Testing Materials, Vol. 34, 2005, pp. 557572.
Sunder, R., Prakash, R. V., and Mitchenko, E. I., Growth of Artically and Naturally Initiating Notch Root Cracks under FALSTAFF Spectrum Loading, Report
No. 797, AGARD, Paper 10, 1994.
Sunder, R., Prakash, R. V., and Mitchenko, E. I., Fractographic Study of Notch Fatigue Crack Closure and Growth Rates, ASTM Spec. Tech. Publ., J. E. Masters and
L. N. Gilbertson, Eds., Vol. 1203, 1993, pp. 113131.
Lang, M., A Model for Fatigue Crack Growth, Part I: Phenomenology, Fatigue
Fract. Eng. Mater. Struct., Vol. 23, No. 7, 2000, pp. 587601.
Lados, D. A., Apelian, D., and Donald, J. K., Fracture Mechanics Analysis for Residual Stress and Crack Closure Corrections, Int. J. Fatigue, Vol. 29, 2007, pp. 687
694.
Seshadri, B. R. and Newman, Jr., J. C., Elastic-Plastic Finite Element Contact
Stress Analyses of Tapered Fasteners, 4th Joint DoD/FAA/NASA Conference on
Aging Aircraft, St. Louis, MO, May 2000.
Seshadri, B. R. and Newman Jr., J. C., Numerical Investigation of Interference-Fit
Tapered Fasteners, USAF Aircraft Structural Integrity Program Conference, San
Antonio, TX, Dec 2000.
Ro, Y., Agnew. S. R., Bray, G. H., and Gangloff, R. P., Environment-Exposure
Dependent Fatigue Crack Growth Kinetics for Al-Cu-Mg/Li, Mater. Sci. Eng., A
Vol. 468470, 2007, pp. 8897.

Potrebbero piacerti anche