You are on page 1of 27

Engineering Fracture Mechanics 208 (2019) 45–71

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Review

Review on failure behaviors of fusion welded high-strength Al


T
alloys due to fine equiaxed zone
Y.N. Hua, S.C. Wua, , L. Chenb

a
State Key Laboratory of Traction Power, Southwest Jiaotong University, Chengdu 610031, China
b
AVIC Manufacturing Technology Institute, Beijing 100024, China

ARTICLE INFO ABSTRACT

Keywords: Fine eQuiaxed Zone (FQZ) close to the fusion line is a common and unique microstructural
Aluminum-lithium alloys feature for fusion welded high-strength Al alloys particularly when containing Zr element. It is
High-strength Al alloys well believed that the presence of FQZ can degrade the fatigue cracking resistance and tensile
Nondendritic equiaxed zone strength of the joints. Thus, a thorough knowledge on the formation and microstructure due to
Fatigue crack initiation and propagation
FQZ and resultant softening and failure mechanisms is necessary to be well understood on the
Synchrotron radiation X-ray microtomography
damage tolerance and structural integrity of fusion welded joints. Current paper reviews the
relevant aspects mentioned above as an important reference to be capable of correlating the
microstructural response with service performance. Some potential solutions to coordinate the
FQZ are therefore discussed. Furthermore, few papers up to date have focused on the fatigue and
fracture mechanisms related with FQZ inside high-strength Al alloy fusion welds. By combining
conventional characterization and numerical modelling approach, the high-resolution synchro-
tron radiation X-ray computed microtomography shows great potential in revealing the com-
petition mechanism between FQZ and welding defects on the cracking behavior.

1. Introduction

Reducing weight and improving damage tolerance have become the most kernel element of structural design and manufacturing
of the aviation, military ship and railway vehicle industries. This is probably due to urgent considerations of payload capability
increase, flight mileage extension, fuel consumption reduction and operation cost saving [1–4]. Thus, those excellent lightweight
materials with synergetic mechanical and fatigue properties, as like high-strength Al-Li alloys, have become one favorable option.
Lithium is thought to be the lightest metal element. Newly-developed Al-Li alloy has been well regarded as the desirable light-
weight material in weight-critical and stiffness-critical structures, mainly due to its attractive natures of low density, high specific
strength and stiffness, excellent stress corrosion and fatigue resistance [1–7]. It is found that adding 1% lithium to Al matrix can
induce a reduction in density of approximately 3% and an increase in elastic modulus of 6% or so [1,8]. The use of high-strength Al-Li
alloys instead of conventional Al alloys can reduce the structural mass by at least 10% and increase the rigidity by up to 15% [1,8].
Therefore, Al-Li alloy has become a desirable and potential substitute for commercial 2000 and 7000 series high-strength Al alloys,
and considerable efforts have been made to extend such community [1,2,9,10].
The first Al alloy with less Li element, called the Scleron alloy, was developed in the early 1920s. The initial commercial 2020 Al-
Cu-Li alloy was subsequently invented by Alcoa of United States in the late 1950s and was extensively applied into the wing surface
and tail of the RA-5C Vigilant aircraft. However, the commercial use of 2020 alloy only lasted for a short period of time because of


Corresponding author.
E-mail address: wusc@swjtu.edu.cn (S.C. Wu).

https://doi.org/10.1016/j.engfracmech.2019.01.013
Received 6 July 2018; Received in revised form 20 November 2018; Accepted 8 January 2019
Available online 09 January 2019
0013-7944/ © 2019 Elsevier Ltd. All rights reserved.
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

low ductility and poor toughness. In the 1960s, the first type of Al-Mg alloy with Li (142× alloy) was developed in the former Soviet
Union and 1420 alloy was considered as one of the most mature Al-Li alloys. In 1989, the Soviet Union successively invented 1430,
1440 and 1450 series alloys. Afterwards, one type of high-strength 1460 Al-Li alloy with the addition of Sc was developed in Russia
through calibrating Cu and Li contents within Al-Li-Cu-Zr alloy [1,11].
In the early 1980s, 8090 and 2091 alloys were introduced in western European as the substitutes for high-strength 2014 and
2024, and 2090 alloy was developed in the United States. The design goal of second-generation Al-Li alloys focused on further
reducing density and increasing stiffness by adding 2 wt% Li. Nevertheless, the addition of Li should be carefully controlled because
the mechanical strength strongly depends on Li content and exhibits the erratic behavior beyond a certain level. It is found that the
satisfactory mechanical property can be obtained in the range of 1.1–1.3 wt% Li. Thus, adding 2 wt% or more lithium into Al matrix
necessarily produces low ductility and toughness, poor corrosion resistance and serious microstructure anisotropy [1,8,11].
In the late 1980s, the development of Al-Li alloys was focused on improving strength and toughness as well as weldability. Some
novel Al-Cu-Li alloys have been invented with unique advantages of high toughness (e.g. 2097 and 2197), excellent cracking re-
sistance (e.g. C-155) and favorable weldability (e.g. Weldalite series) over standard alloys. For such Al-Cu-Li alloys, the content of Li
was further decreased and the minor addition of Zr, Mg, Zn, Ag and rare earth metals could achieve a satisfactory combination of
strength and toughness, such as 2050, 2060, 2099, 2195, 2196, 2198, 2199, 2297 and 2397 alloys. The third-generation Al-Li alloy
was thus able to compete with traditional high-strength 2000 and 7000 series Al alloys [1,11,12].
It is well-known that riveting has been the main joining technology of Al alloy structural components such as aircraft body and
wings for a very long time [1]. However, riveting is actually a very tedious operation with slow assembling and especially the
limitation when joining thin elements [10,13]. As an epochal occurrence, to effectively join large-scale lightweight structures, various
welding methods have been developed in the last decades, particularly on advanced high energy density beam welding (mainly the
laser and its hybrid with arc) and solid-state friction stir welding (FSW) except for conventional arc welding of metal inert-gas (MIG)
and tungsten inert-gas (TIG) [14–17]. Recently, Al-Li alloy welded structures have gained more and more increasing interests in
scientific aspects and engineering applications.
High energy density beam welding technologies such as laser beam welding (LBM) and electron beam welding (EBM) are well
believed to be the promising joining methods for high-strength Al alloy thin plates. These advanced welding procedures can achieve a
high-quality welded joint with the high welding efficiency, narrow and deep weld, small heat-affected zone (HAZ) and low thermal
deformation. Furthermore, compared with EBM limited in a vacuum cell, LBM is proved to be a powerful tool to produce large
complicated structures in terms of the excellent flexibility and automation [10,13,18,19]. It is reported that laser beam welding has
been successfully applied as an alternative to riveting in manufacturing Al alloy lower fuselage panels in Airbus [1].
Extensive investigations focus on the development of novel Al-Li alloys. Critical to the successful application of these high-
strength Al-Li alloys will be the weldability, including the gas porosity, weld solidification cracking and HAZ liquation cracking. In
particular, Al-Li alloys are also very sensitive to fusion line cracking, which is closely related with a Fine eQuiaxed Zone (FQZ) along
the fusion line (FL). The narrow FQZ is composed of very fine nondendritic equiaxed grains surrounded by grain boundary eutectic
products, giving the poor mechanical and corrosion properties [20,21]. It is reported that the quasi-static tensile fracture pre-
ferentially proceeds along FQZ within Al-Li alloy butt-welds and T-joints [22–26]. The FQZ behavior has a considerable influence on
the tensile strength and fracture characteristic of the joints [24]. Also, the cracks are found to propagate rapidly to FQZ during
fatigued loading [5]. Thus, a detailed knowledge on the formation mechanism, microstructural feature and failure behavior of FQZ is
particularly revealed to improve damage tolerance and structural integrity of Al-Li alloy fusion welds.
The formation and presence of FQZ was first observed within Al-Li alloy welds in the early 1990s. The well-known fine equiaxed
zone is a common and unique microstructural phenomenon for Al-Li alloy and other high-strength Zr-containing Al alloy fusion
welds, produced by laser beam source, variable polarity plasma arc, electron beam and gas tungsten arc welding [25,27,28]. Here the
high-strength Zr-containing Al alloy mainly refers to the 7000 series Al alloy with Zr element. The formation of FQZ is closely related
to Zr and Li, with Zr exhibiting a more dominant effect [28]. Compared with 7000 series Al alloys, due to the addition of Li, FQZ
within Al-Li alloy welds is more obvious and results in the poor mechanical and corrosion properties. Thus, the examples in this paper
are mainly about Al-Li alloys and a few 7000 series Zr-containing Al alloys.
It should be pointed out that there also exists a fine equiaxed zone within steel welded joints. However, its distribution feature and
formation mechanism are quite different from the FQZ inside Al alloy fusion welds. As to steel welds, the fine equiaxed zone is located
in the HAZ, generally referred to as fine-grained heat-affected zone (FGHAZ). FGHAZ is far from the fusion line and the peak
temperature is slightly higher than Ac3, leading to the dissolution of some carbides and limited grain growth. The carbide coarsening
and microstructure deterioration can weaken the dispersion/precipitation strengthening effect, giving degraded hardness and creep
strength of FGHAZ. Due to the higher creep strain rate and lower creep strength, FGHAZ is regarded as major IV type cracking sites.
Here, we give a brief introduction of FGHAZ, on the one hand, reminding to note to distinguish FQZ within Al alloy fusion welds from
FGHAZ within steel welds. On the other hand, the study of FGHAZ can serve as a basic reference for investigating the FQZ [29,30].
This work focuses on the recent progress related with FQZ of various high-strength Al alloy fusion welds produced by arc, laser
and their hybrid. Such unique characteristic near the HAZ has been usually believed to behave the harmful effect on mechanical and
fracture properties. To fully develop the capacity of high-strength Al alloys, it is therefore valuable to explore its nature and resultant
influence at the presence of the FQZ. Initially, the formation of FQZ and its influence factors such as the chemical composition,
welding parameter, base metal (BM), formation temperature and molten pool turbulence are introduced. Subsequently, micro-
structural features of FQZ including the microstructure, grain boundary segregation and solidification cracking are discussed. Then
the softening behavior, quasi-static tensile and impact properties together with the fatigue and fracture resistance are highlighted.
Besides, some potential solutions to calibrate the FQZ from Al-Li alloy welds are also compared. Currently, very few papers or

46
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 1. The microstructures of Al alloy fusion welds: (a) the schematic diagram of conventional Al alloy welds (marked in green) and Al-Li alloy
welds (marked in yellow) and (b) the microstructure evolution related with solidification process within conventional Al alloy fusion welds. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

investigations have focused on the fatigue and fracture behaviors due to FQZ mainly because of its small size in width. More fatigue
life data and theoretical aspects as well as the prediction model and numerical simulation are expected to be performed in near
future.

2. Formation mechanism

Al alloy fusion welds present various microstructures near the FL within fusion zone (FZ). The FL can be actually considered as an
interface between the FZ and partially melted zone (PMZ). Generally, the weld pool solidification begins with the epitaxial growth
from partially melted grains due to excellent nucleation sites. The grain growth rate is far greater than the nucleation rate, thus
resulting in the long and oriented columnar crystal structures close to the PMZ [9,22,25,31,32], as shown in Fig. 1(a) marked in
green.
However, as to Al alloys containing Zr and Li, the formation and presence of FQZ within FZ prevent grain epitaxial growth from
PMZ, as illustrated in Fig. 1(a) marked in yellow. FQZ mainly consists of near spherical equiaxed grains with random crystal or-
ientation and the average grain diameter is merely 3–6 μm [22,25,33]. The nondendritic morphology of FQZ apparently deviates
from typical microstructures that can be explained by the constitutional undercooling theory, such as planar, cellular, dendritic or
equiaxed dendritic structures, as shown in Fig. 1(b) [22]. Thus, the formation mechanism of FQZ has gained increasing attentions in
the metallurgical community.
Two hypotheses have been put forward about the formation mechanism of FQZ. The first hypothesis proposed by Shah in 1992
was that FQZ was located in the partially melted zone and formed via recrystallization [34]. The second one put forward by Gutierrez
and Lippold was based on the heterogeneous nucleation theory within a stagnant liquid layer close to the FL, which was defined as
the unmixed zone (UMZ) [22,35]. To determine the formation mechanism, a systematic investigation has been performed in terms of
the alloying composition, solidification substrate and welding parameter [21,22,25,28].
The formation process of FQZ is proved to be closely related to Zr and Li, with Zr exhibiting a more dominant effect [28]. Reddy
et al. employed two kinds of BMs to reveal the alloying element effect on the FQZ [25]. One was 8090 Al-Li-Cu alloy with the addition
of 0.1 wt% Zr and the other alloy with similar compositions to 8090 but without Zr was designated by 8090 (-Zr). Fig. 2 presents the
optical microstructure close to the FL inside TIG welded 8090 and 8090 (-Zr) alloys [25]. A 250-μm-width FQZ is distinctly found

47
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 2. The microstructure near the FL in TIG welded (a) 8090 and (b) 8090 (-Zr) [25].

close to the FL within 8090 TIG welds in Fig. 2(a). While concerning 8090 (-Zr) TIG welds, no FQZ can be seen and typical epitaxial
growth is observed near the FL (see Fig. 2(b)).
FQZ can also be found in Zr-containing Al alloys without the addition of Li, such as high-strength 7000 series Al alloys including
7020, 7050, 7 N01, 7018 and RDE40, which confirms that FQZ is not unique to Al-Li alloy welds [15,21,22,36]. A 50-μm-width FQZ
is found within 7020-T651 and 7050-T7451 hybrid laser welds, as shown in Fig. 3 [15]. It is clear that Zr plays a crucial role in the
formation of FQZ. Note that no FQZ is observed under very low levels of Zr and Li, such as 0.03 wt% Zr and 0.5 wt% Li [28].
In addition to Zr and Li inside original base material, elements of the filler metal also have a considerable effect on the FQZ. It is
found that the FQZ width increases with an increase of Zr and Li in the filler metal [9,21]. Lin et al. employed newly-developed filler
metals with different contents of Zr and Li to illustrate the effect of filler metal on the FQZ, as shown in Fig. 4 [9]. The width of FQZ is
104 μm and 180 μm in TIG welded 2090 Al-Li-Cu alloys, with the addition of 0.4 wt% Zr and 0.93 wt% Zr into the filler wire
respectively. Moreover, it is found that calibrating Zr-containing Al alloys via adding minor Er and Sc can facilitate the formation of
FQZ mainly due to the similar Al3M (M may be Zr, Er, Sc, LixZr1−x, ErxZr1−x) L12-ordered crystal structure during the solidification
[21,37–39].
The solidification substrate also appears to play a crucial role in the formation of FQZ. It is found that no FQZ can be observed in
the as-cast and as-welded substrate conditions. However, the FQZ will appear if the BM is in the solution treated and aged, annealed
or other homogenized conditions [25,28]. Fig. 5 presents the optical microstructure near the FL within 8090 Al-Li-Cu alloy welds,
with different solidification substrates of as-cast together with solution treated and aged [25]. The FQZ is clearly observed in TIG
welded 8090 alloys with the base material of being solution treated at 535 °C for 45 min and aged at 170 °C for 24 h (see Fig. 5(b)).
It has been further observed via the transmission electron microscopy (TEM) that fine Al3Zr particles of approximately 20–40 nm
in diameter are well located within small-sized equiaxed grains of FQZ, as shown in Fig. 6 [23]. In contrast, no Al3Zr can be found in
areas where normal dendritic solidification occurs. The analysis shows that FQZ is formed only when the BM contains element Zr and
is homogenized or further processed to precipitate the phase Al3Zr. Besides numerous experimental analyses, the microstructure
evolution close to the FL was simulated by Kostrivas and Lippold based on the Gleeble® 1500 thermo-mechanical simulator, which
tried to show the effects of chemical composition and BM nature on the FQZ [22]. The predicted results well coincided with ex-
perimental analyses in terms of FL microstructures.
As to welding parameter, it is found that process parameters are significantly related with the FQZ width but show relatively very
less effect on the formation of FQZ. The width of FQZ usually increases with the reduction of welding heat input. High energy density
beam welding methods such as LBW and EBW can generally facilitate the formation of FQZ due to high welding speed and low heat
input [5]. However, it is not possible to completely eliminate FQZ even over a wide range of heat input achieved by varying the

Fig. 3. The microstructure near the FL in hybrid laser welded (a) 7020-T651 and (b) 7050-T7451 [15].

48
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 4. The microstructure near the FL in TIG welded 2090 alloys with the addition of (a) 0.4 wt% Zr and (b) 0.93 wt% Zr into the filler wire [9].

Fig. 5. The microstructure near the FL within TIG welded 8090 alloys with different solidification substrates of (a) as-cast and (b) solution treated
and aged conditions [25].

Fig. 6. The FQZ precipitate characteristics in laser welded 2060-T8/2099-T83 Al-Li-Cu alloys: (a) high-resolution TEM image of spherical particles
within grains, (b) selected area diffraction pattern taken from the blue frame in (a) along Al matrix [0 0 1] and (c) schematic selected area diffraction
pattern of (b) [23]. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

travelling speed and arc current during autogenous gas tungsten arc welding, although the width and grain size of FQZ vary con-
siderably [28].
Based on the comprehensive investigations about the effects of alloying composition, solidification substrate and welding
parameter on the formation of FQZ, it is now believed that the heterogeneous nucleation due to Zr-containing precipitates, such as
Al3Zr and Al3(Lix,Zr1−x) inside original BM, should be the well-accepted formation mechanism of FQZ, rather than previously
proposed recrystallization behavior in the PMZ [34]. The nucleation and growth of small-sized nondendritic equiaxed grains within
FQZ will be discussed in detail as below.
It is well-known that the partition coefficient of Zr is greater than 1 in Al-Zr phase diagram, indicating that Zr will be kept in the
intragranular solid solution during the rapid solidification [21,22]. Enough equilibrium Al3Zr particles are expected to precipitate

49
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 7. Dark field TEM micrographs of the particle Al3Zr with (a) L12 and (b) DO23 structures [33].

after aging. The melting temperature of Al3Zr is generally as high as 1055 °C [39]. Also, the thermally stable intermetallic Al3Zr has
an L12 ordered face-centered cubic structure with the lattice parameter 0.409 nm, very close to the Al matrix 0.405 nm (face-centered
cubic structure, FCC) [22,37,39–41]. As a result, these equilibrium phases are expected to provide sufficient and effective hetero-
geneous nucleation sites during the solidification. Moreover, since there is not enough time and space available for grains to grow
branches before contacting with each other, a large number of small-sized equiaxed grains with near sphericity are formed within
FQZ.
However, it is found that the crystal structure modification of Al3Zr occurs from L12 ordered face-centered cubic structure to DO23
tetragonal structure when the temperature exceeds 600 °C [22,33,42], as shown in Fig. 7. The DO23 tetragonal structure will result in
serious lattice mismatch between Al3(Lix,Zr1−x) and α-Al matrix. Therefore, it is concluded that both the number of effective het-
erogeneous nucleation sites and nucleation efficiency are necessarily reduced as the peak temperature increases. The heterogeneous
nucleation behavior only happens to a small amount of grains within FQZ during the solidification. In view of dispersed nucleation
sites, it is easily deduced that these grains will have sufficient time and space to grow and develop branches before contacting with
neighboring grains, giving dendritic structures. Besides, due to limited heterogeneous nucleation sites, typical epitaxial growth from
partially melted grains may occur in the FQZ.
Kostrivas and Lippold confirmed that FQZ was confined in the temperature range of 630–640 °C and dendritic structures were
formed when the temperature was above 640 °C (see Fig. 8) [22]. In summary, it is believed that heterogeneous nucleation from Al3Zr
and Al3(Lix,Zr1−x) inside original BM in a narrow temperature range leads to the formation of FQZ. A thin liquid layer with similar
alloying elements to the BM exists along the liquid/solid FL and the temperature is close to the melting point of BM, isolating from the
high temperature in the weld center [9,33]. The moderate temperature and fluid flow condition at the FL are conducive such that
particles neither dissolve nor sweep into the weld, contributing to the formation of FQZ [9]. The higher temperature towards weld
center leads to considerable dissolution of Al3Zr and consequently the columnar crystal zone forms due to the epitaxial growth from
small-sized grains within FQZ.
For Al-Li alloy welds, the element Li also plays an important role in the formation of FQZ [9,10,33,34,43]. First, the addition of Li
can effectively increase the volume fraction of precipitates by decreasing Zr solid solubility in aluminum. Then, the higher tem-
perature phase Al3(Lix,Zr1−x) are able to form by replacing the atom Zr of Al3Zr by Li. Compared with the particle Al3Zr, these new

Fig. 8. The FL microstructure within 2195-T8 Al-Li-Cu alloys produced using Gleeble® with the peak temperature of (a) 630 °C and (b) 660 °C [22].

50
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 9. The FQZ in the FZ inside TIG welded 8090 alloys under (a) pulsed current and (b) arc oscillation conditions [25].

ternary particles are expected to provide more potential heterogeneous nucleation sites and have higher nucleation efficiency since
the lattice mismatch with α-Al matrix is much lower. Besides, lithium is a surface-active element that can easily absorb onto the
surface of Al3Zr and Al3(Lix,Zr1−x). It is beneficial to decrease the interfacial energy between precipitates and growing solids, making
great contribution to heterogeneous nucleation. And Li is considered as the most effective element to reduce the surface tension of
liquid aluminum [43]. As a result, lithium can generally facilitate the formation of FQZ in Al-Li alloy welds with fair finer equiaxed
grains based on the classical heterogeneous nucleation theory as below [43]:

16 3LS 2 3 cos + cos3


G=
3( G V ) 2 4 (1)

where ΔG is the energy barrier of heterogeneous nucleation, σLS is the liquid–solid interfacial tension, ΔGV is the Gibbs free energy
difference between liquid and solid phase per unit volume, and θ is the contact angle.
The discussions above primarily deal with the continuous current (CC) weld and it is found that FQZ is mainly located close to the
FL in CC Al-Li welds. While as to pulsed current (PC) and arc oscillating (AO) Al-Li welds, FQZ is not restricted to the FL but in the FZ
[21,25]. As is shown in Fig. 9 that FQZ is well located in the weld center inside TIG welded 8090 Al-Li-Cu alloys under pulsed current
and arc oscillation conditions [25]. The FZ presents alternate microstructure bands of nondendritic equiaxed grains and dendritic or
columnar crystal structures. The FQZ within PC welds in Fig. 9(a) appears to be located at curved boundaries along the trailing edge
of molten pool. For AO welds, fine equiaxed grains over a wider zone are found to be alternately arranged with coarse dendritic
structures, as shown in Fig. 9(b).
It has been confirmed that FQZ is mainly confined to the FL within CC welds due to favorable temperature and fluid flow
condition [9,22]. Concerning PC and AO welds, the formation and presence of FQZ in the FZ are attributed to fine equiaxed grains
transferring from FL to the weld center driven by fluid flow [21,25,44,45]. Different fluid flow modes such as radial flow and
rotational flow have been pointed out during PC welding [25]. Only chill crystals near the trailing edge of weld pool are expected to
survive owing to the local lower temperature, thus giving the curved FQZ band. Besides, the periodic current variation can necessarily
enhance the convective force, leading to the chemical composition homogenization and element segregation reduction, as well as
decreasing the temperature gradient and peak temperature. The fluid flow is more complex under the AO condition as a result of
constantly changing weld pool orientation, giving the strong alternate distribution of nondendritic and dendritic structures.

3. Microstructural features

3.1. Microstructure

The service performance is well-known to correlate with microstructural features [46]. A clear understanding of the micro-
structure, element segregation and solidification cracking of FQZ is thus expected to better reveal the mechanical and fatigue per-
formance. Fig. 10 presents the crystal morphology and orientation as well as grain size inside 2A97 Al-Li-Cu laser butt welds via
optical microscopy (OM) and electron backscattered diffraction (EBSD) [47]. The FQZ is well located close to the solid-liquid FL,
preventing the conventional epitaxial growth from partially melted grains.
It is observed from Fig. 10(b) that FQZ with about 50 μm in width is composed of large number of near spherical and nondendritic
fine grains. Furthermore, these small-sized grains present random orientation (see Fig. 10(c)), deviating from the preferred crystal
orientation 〈 0 0〉 of BM [5]. It is believed that the lack of any preferred crystallographic orientations further supports the hetero-
geneous nucleation theory proposed by Gutierrez and Lippold [20,21]. Besides, it is worth noting that the acute grain size gradient
exists on both sides of such a narrow FQZ [48]. And the nondendritic equiaxed grains of FQZ are separated from partially melted
grains by the high angle grain boundaries (> 15°), giving a weak transition zone [28,49]. The grain equivalent diameter in Fig. 10(d)
is less than 9 μm and the average grain size is merely 3.7 μm inside 2A97 laser welds. In general, the FQZ grain size within all
investigated Al-Li alloy welds varies from 3 μm to 10 μm [27].
It is typically observed that FQZ is absent at the joint top while the width gradually increases to maximum at the bottom [13,22].

51
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 10. The microstructure in laser welded 2A97 alloys: (a) OM based microstructure across the joint, (b) OM based microstructure near the FL, (c)
EBSD based microstructure close to the FL and (d) the grain size distribution within FQZ [47].

The inhomogeneity in width can be attributed to different fluid flow conditions along the thickness direction [50]. It is known that
thermally stable intermetallic particles including Al3Zr and Al3(Lix,Zr1−x) are able to survive under appropriate temperature within
the stagnant liquid layer, and thus fine nondendritic equiaxed grains are formed via heterogeneous nucleation during the solidifi-
cation. However, considering the fluid flow characteristic, some intermetallic particles are swept into the weld center through
complicated liquid flowing, leading to particles dissolution and resultant less nucleation sites. The fluid flow pattern is expected to be
vigorous near the joint top, giving a quite narrow FQZ. By contrast, a wider FQZ is generated at the weld bottom mainly due to
relatively weak molten pool disturbance [22].
In addition to Al-Li alloy butt welds, FQZ also exists inside other types of welded joints. Fig. 11 shows the distribution char-
acteristic and width of FQZ along the lower and upper fusion lines inside double-sided laser welded 2060-T8/2099-T83 T-joints
[23,24]. The width of FQZ varies significantly from 5 μm to 100 μm. It is demonstrated that uneven temperature and inconsistent
velocity are the two key factors for the width inhomogeneity inside double-sided LBW T-joints [23]. It is found by FLUENT simulation
that not only the lowest temperature but also the minimum velocity are located along the upper and lower fusion lines, and parti-
cularly concentrated around bottoms of the two fusion lines [50,51]. Therefore, more thermally stable intermetallic particles can be
reserved as enough and effective heterogeneous nucleation sites for fine equiaxed grains.

3.2. Element segregation

The grain boundary segregation and intra-grain precipitation are common in Al alloy welds [5]. Fig. 12 shows the FL micro-
structure within laser welded 2A97 Al-Li-Cu alloys via scanning electron microscope (SEM) [10]. The brighter color indicates element
enrichment with relatively higher atomic weight compared to the darker gray for α-Al matrix. As is shown in Fig. 12 that grain
boundaries are decorated with eutectic constituents and second particles due to element redistribution under non-equilibrium
crystallization. It is demonstrated that the segregation of solute element as well as the presence of liquid film at the grain boundaries
generally lead to the poor intergranular mechanical and corrosion properties in the FQZ [21].
Concerning eutectic products along the grain boundaries, it is found that the difference in element segregation can necessarily
change the eutectic morphology, which is closely related with the mechanical property of the joints [13,33]. Dev et al. investigated

52
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 11. The microstructure and width of FQZ in double-sided laser welded 2060-T8/2099-T83 Al-Li-Cu alloy T-joints: (a) the microstructure across
the weld, especially showing the FQZ close to the lower and upper fusion lines, and (b) and (c) the width of FQZ along the lower and upper fusion
lines, respectively [23,24].

Fig. 12. Backscattered SEM based FL microstructure of laser welded 2A97 alloys [10].

the intergranular corrosion behavior of FQZ according to ASTM G 110–92 and tried to reveal the relationship among the solute
segregation, eutectic morphology and intergranular corrosion property [21]. Fig. 13 presents the grain boundary attack of FQZ inside
AA7018 (0.12 wt% Zr) and RDE40 (0.18 wt% Zr) 7000 series alloy TIG welds.
It is clear that eutectic constituents along the grain boundaries within AA7018 FQZ are found to be continuous, while the eutectic
products are relatively less and discontinuous within RDE40 FQZ. The severe grain boundary attack and continuous eutectic

53
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 13. The grain boundary attack of FQZ inside (a) AA7018 and (b) RDE40 TIG welds [21].

morphology of AA7018 FQZ may attribute to the larger potential segregation of Cu and Zn. Actually, RDE40 belongs to Al-Zn-Mg
alloy with no Cu and low Zn/Mg ratio. By contrast, AA7018 has relatively higher Cu and Zn/Mg ratio.
In addition to the difference in solute segregation and eutectic morphology among different Al alloy welds, the eutectic structure
also varies from FQZ to the weld center in the same Al alloy weld. Fig. 14 shows the FL microstructure of laser welded 2060 Al-Li-Cu
alloys based on the backscattered SEM [13]. It can be observed that eutectic products exist along the grain boundaries of FZ, FQZ and
HAZ due to the solute segregation under non-equilibrium solidification. As to eutectic morphology, it is found that the continuous
network eutectic product is formed within FQZ, while it becomes rather unconnected in the columnar crystal zone [13].
By using electron probe microanalyzer (EPMA), Han et al. investigated the element distribution of Si and Cu near the lower fusion
line inside 2060/2099 double-sided laser welded T-joints with the filler AA4047 (AlSi12) [23], as shown in Fig. 15. Since the Si
content in the base material is extremely low, almost all of Si in the weld comes from the filler metal. It is found that as the width of
FQZ increases to about 50 μm, the content of Si within FQZ decreases dramatically. Actually, Si is mainly concentrated in the local
FQZ very close to the columnar crystal zone and almost no Si is detected in the local FQZ near the PMZ (see Fig. 15(a)). It is believed
that the diffusion of Si in a wider FQZ is primarily suppressed by the laminar boundary layer along the fusion line.
However, the distribution of Cu within FQZ is not as clear as Si, but the Cu content in FQZ could be higher compared to FZ due to
the higher brightness in Fig. 15(b). It is believed that with respect to HAZ, PMZ and FZ, the higher content of Cu within FQZ is
probably attributed to the volatilization of Li, Zn and Mn [23]. The Cu content of FZ decreases sharply under the dilution of filler
wire. While, the dilution effect of filler metal is greatly suppressed in FQZ due to the laminar boundary layer along the fusion line.
Thus, the difference in alloying element between FQZ and other zones of the joints is obvious, necessarily resulting in distinct eutectic
structures and hot cracking tendency.
Besides the FQZ eutectic constituents inside continuous current welds, the eutectic morphology along the grain boundaries is also
investigated in the pulsed current and arc oscillating welds [21], as illustrated in Fig. 16. ASTM G 110-92 test clearly presents the
continuous grain boundary attack in the case of CC FQZ whereas more divorced grain boundary attack with regard to PC and AO. The
intergranular corrosion cracking resistance of CC FQZ is thus lower compared with PC and AO. The less and discontinuous eutectic
products at the grain boundaries within PC and AO welds are primarily attributed to the lower solute segregation of Zn and Mg. It is
well believed that the enhanced fluid disturbance in the weld pool is beneficial to homogenize the alloying composition and then
reduce the solute segregation.

Fig. 14. The FL microstructure in laser welded 2060 alloys: (a) the optical microstructure across the FL and (b) detail of the region marked by a
rectangle in (a) based on the backscattered SEM [13].

54
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 15. The element planar distribution near the lower fusion line in 2060/2099 double-sided laser welded T-joints with the filler AA4047: (a)
element Si and (b) element Cu [23].

Fig. 16. The grain boundary attack of FQZ in TIG welded RDE40 under (a) continuous current, (b) pulsed current and (c) arc oscillation conditions
[21].

3.3. Solidification cracking

It is known that Al-Li alloy fusion welds have the high tendency to solidification cracking [26]. It is demonstrated that the
solidification cracking of AI-Li-X alloy welds is often associated with FQZ [22]. The nondendritic equiaxed grains within FQZ are
surrounded by grain boundary eutectic constituents [20]. The eutectic products due to the element segregation, and liquid films at
the grain boundaries generally result in the poor resistance to intergranular cracking [21]. A clear understanding on the formation of
hot cracks within FQZ is urgent. Fig. 17 presents the typical hot cracks of FQZ near the upper fusion line inside double-sided laser
welded Al-Li alloy T-joints [24].
Han et al. carried out a detailed investigation on the relationship between the distribution of alloying elements and formation of

Fig. 17. The FL microstructure in double-sided laser welded Al-Li alloy T-joints: (a) cross-section diagram and (b) hot cracks within FQZ near the
upper fusion line [24].

55
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 18. EDS analysis of Si and Cu across the lower fusion line in 2060/2099 double-sided laser welded T-joints: (a) FL microstructure and (b)
element distributions of Si and Cu along the solid line in (a) [23].

intergranular microcracks within FQZ inside 2060/2099 double-sided laser welded T-joints with the filler AA4047 (AlSi12), as shown
in Fig. 18 [23]. The element distribution on the solid line close to the microcracks is quantitively characterized via the energy
dispersive spectroscopy (EDS). It is found that the content of Si and Cu within FQZ are lower and higher compared to HAZ, PMZ and
FZ, respectively. It is generally believed that Cu is a major element in the formation of intergranular cracks, while Si has a strong
function of repairing solidification cracking [52–54]. Thus, the inhomogeneous distribution of Si and Cu plays a crucial role in the
intergranular microcracks of FQZ.
It is well-confirmed that the diffusion of solute element to intergranular eutectic can result in considerable intergranular seg-
regation and reduce the eutectic melting point [23]. As the temperature of FQZ drops below the solid phase line, the contraction
deformation of equiaxed grains is believed to induce the intergranular tensile stress within eutectic constituents. However, due to the
local absence of Si in the wider FQZ, Si eutectic products are not sufficient to refill the intergranular cavity induced by the residual
tensile stress and low melting point divorced eutectics. Consequently, intergranular microcracks will be formed in the FQZ.

4. Mechanical behavior

4.1. Softening phenomenon

Softening behavior is quite common inside high-strength Al alloy fusion welds [14,46,55]. It is well-known that four types of
strengthening mechanisms have been established for Al alloys, including the precipitation strengthening, solid solution strength-
ening, grain boundary strengthening and dislocation strengthening [56,57]. Based on these theories, Wu et al. conducted a detailed
investigation on the hardness distribution across fusion welded AA7075-T6 without Zr, as shown in Fig. 19 [46]. It is found from
Fig. 19(a) that weld center is the weakest part of the joints, giving the lowest hardness. The softening degree is defined as the ratio of
the hardness of weld center and BM (≈157 HV), and then the softening level is 57.3%, 66.8% and 71.3% for the MIG, hybrid and
laser welded joints, respectively. It is believed that the hardness loss of FZ mainly results from two aspects: the evaporation loss of
strengthening elements and modification of precipitates [14,46,58].
However, for Al-Li alloy and other Zr-containing Al alloy welds, due to the formation and presence of FQZ, what is the hardness
distribution characteristic across the joints? Fu et al. and Ning et al. found that the lowest microhardness in laser welded Al-Li alloy
butt welds was located in the weld center [5,10]. Huang et al. proposed that the columnar crystal zone was the weakest part of the
joints, giving the lowest microhardness [59]. These hardness distributions across the joints based on microhardness testing inevitably
ignore the very narrow FQZ. The hardness deviation or incorrect measurement may lead to the wrong research conclusion and have a
negative influence on the safety assessment of welded structures. Thus, how to accurately characterize the hardness of such a narrow
FQZ is vital to reveal the softening behavior of Al-Li alloy welds.
The nano-indentation with much smaller indenter tip and loading force is considered as a powerful tool to accurately identify the
hardness distribution especially for FQZ [23]. Wu et al. performed an investigation on the nanohardness distribution across AA7050
hybrid laser welds via the KEYSIGHT G200 Nano Indenter with a commercially available diamond Berkovich tip, as shown in
Fig. 20(a). Displacement-controlled indentation test was conducted with the maximum depth limit of 1000 nm and the average value
of curve flat section (750–900 nm) was taken as the hardness of each measured point. It is clear from Fig. 20(a) that FQZ gives the
lowest hardness and the softening degree is 0.5 or so. By contrast, the weld center is the weakest part of the same joint via classical
microhardness testing (see Fig. 20(b)). The abnormal hardness deviation from nanohardness measurement further confirms the
necessity of nano-indentation technology in accurately characterizing the hardness of FQZ.
The softening behavior of Al-Li alloy fusion welds is different from that of conventional Al alloy welds. An important point is
found that FQZ presents the lowest hardness of the joints, such as the hybrid laser welded 7050 alloys, laser welded 2A97 and 2060
Al-Li alloys and TIG welded 2219 alloys in Fig. 21 [13,47,60–62]. The softening behavior of FQZ has gained much attention and

56
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 19. The softening behavior of AA7075-T6 fusion welds: (a) measured mean microhardness distribution across the MIG, hybrid and laser welds,
(b) and (c) the planar scanning of Zn and Cu across a hybrid laser weld, (d) the line scanning of Mg within half hybrid weld and (e)–(g) strengthening
precipitates at the far-field BM, HAZ and weld center, respectively [46].

interest in scientific research. In addition to serious softening behavior, note that both sides of FQZ with acute microstructure
gradient result in the sharp hardness gradient, which are generally regarded as the severe stress concentration for cracking sites
[15,63]. It is of great importance to conduct a comprehensive analysis on the softening behavior of FQZ.
Concerning the grain strengthening, it is known that the contribution of grain boundary to the strength is inversely proportional
to the square root of grain diameter in terms of classic Hall-Petch equation [64]. It is found in AA7050 hybrid laser welds that the
grain size of BM, weld center and FQZ is 31 μm, 56 μm and merely 3 μm, respectively. The highest yield strength and hardness are
expected at FQZ while the macroscopic property (see Fig. 21(a)) gives a contrary result. It is evidently indicated that the grain size
and distribution are less responsible for the strengthening mechanism of FQZ. Moreover, as to precipitation-hardened Al alloys, the
intrinsic parameter k in the Hall-Petch equation is relatively small, which physically suggests the grain size evolution has a very
limited effect on the hardness and strength modification [64–66].
It is well-confirmed that insufficient aging contributes a lot to the softening behavior of FZ within Al-Li alloy welds, certainly
containing the FQZ. Generally, FZ insufficient aging is mainly attributed to the element segregation (already existed in BM or added
by filler wire) at the grain boundaries under extremely non-equilibrium solidification, leading to insufficient supersaturated solid
solution and resultant a few strengthening precipitates. Especially, the precipitates are much less inside the joints subjected to the
high energy density welding due to the high welding speed and limited time to precipitate. Moreover, after being exposed to the high-
intensity heat source, inevitable evaporation loss of low melting point elements such as Zn and Mg will result in fewer strengthening
precipitates without the supplement from filler wires. Thus, the strengthening element redistribution and loss are exactly the primary

57
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 20. Measured mean hardness curves across AA7050 hybrid laser welds with: (a) KEYSIGHT G200 Nano indenter and (b) HVS-30 Vickers
indenter.

Fig. 21. Measured mean hardness curves of (a) 7050 laser hybrid welds, (b) 2A97 laser welds together with (c) and (d) 2060 laser welds [13,47,60].

reason for the degraded hardness and strength.


Kostrivas and Lippold have tried to explain the hardness difference between the weld center and FQZ via the Scheil-Gulliver
equation, as shown in Eq. (2) [22]. The Scheil-Gulliver model, which is applied into a wide range of cooling rates, is valid to assess the
element segregation at the grain boundaries of both equiaxed dendrites and equiaxed non-dendrites [22,67]. It is worth noting that
element segregation is strongly affected by the intergranular liquid flow pattern and the nondendritic equiaxed grains with nearly no
branches will necessarily influence the element segregation and then the eutectic morphology within FQZ, as contrast to equiaxed
dendrites in the weld center. It is believed that the serious element segregation and continuous network eutectic structures along the
grain boundaries are detrimental to the local mechanical behavior.

58
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 22. The local softening behavior in laser welded Al-Li alloy T-joints: (a) the nanohardness distribution across the lower fusion line, (b) and (d)
TEM bright-field micrograph of cross section of indentation L5 and L7, (c) and (e) microstructure evolution schematic diagram under the indentation
L5 and L7 at the maximum loading stage [23].

(CL SS ) fS = (1 - fS ) CL
CS = kC0 (1 fS ) k 1

k 1
CL = C0 fL (2)

where CS and CL are the local composition of solid and liquid, respectively, k = CS/CL is the partition coefficient, C0 is the base
composition, and fS and fL are the fraction of solid and liquid phases, respectively.
Recently, Han et al. confirmed the local softening phenomenon within FQZ via the accurate nano-indentation testing, as shown in
Fig. 22(a) [23]. Both L5 and L6 indentations show the lower nanohardness compared with L4, L7 and L8 of FQZ, and also other
regions of the joints. The combination of TEM and nano-indentation testing is believed to be a powerful tool to reveal the softening
behavior of FQZ, mainly due to the narrow FQZ, fine equiaxed grains and small indents, as shown in Fig. 22(b) and (d). It is directly
observed in Fig. 22(b) and (c) that an intergranular tiny hot crack is formed right underneath the indent tip. Besides, a small amount
of short and thin Al2Cu phases are distributed within equiaxed grains and at their boundaries.
It is believed that intergranular microcracks in the local softening area of FQZ are likely to be caused by the local absence of Si and
corresponding divorced eutectics. And these intergranular microcracks provide almost no resistance to the intergranular plastic
deformation. Moreover, it is known that both intergranular and intragranular strengthening phases play a crucial role in the dis-
location slipping and climbing. However, short and thin Al2Cu phases in the local softening area only provide very limited resistance
to the plastic deformation during indenter loading process. Thus, it is concluded that the local softening phenomenon of FQZ is
primarily attributed to the formation of microcracks and the absence of abundant thick Al2Cu precipitates at grain boundaries.
In summary, grain boundary strengthening is not sufficient to elucidate the hardness distribution of FQZ. The serious element
segregation and resultant insufficient aging play an important role in the hardness loss. The equiaxed grain boundaries are decorated
with eutectic constituents due to the solute segregation. The composition, quantity and morphology of eutectic products contribute a
lot to the hardness variation. It is believed that the continuous grain boundary eutectics or particles result in the relatively poor
strength and plasticity. Thus, from the hardness index that characterizing the comprehensive mechanical properties, severe hardness
loss occurs. Besides, the local softening phenomenon of FQZ is confirmed due to the formation of intergranular microcracks and the
absence of thick intergranular precipitates.

4.2. Impact toughness

In addition to softening behavior, the resistance to notch sensitivity (ductility) of FQZ is also investigated via the Charpy impact

59
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 23. The specimen dimension and notch schematic diagram of Charpy V notch testing.

testing. Fig. 23 shows the specimen dimension and notch schematic diagram of Charpy V notch (CVN) test [21]. Small-sized CVN
specimens are prepared from 6-mm-thick TIG welded RDE40 and AA7018 alloys. The metallographic observation is performed first to
locate the FQZ and make sure the notch is well prepared within FQZ.
The Charpy impact energy values of both TIG welded RDE40 and AA7018 are summarized in the following Table 1 [21]. The
RDE40 welded joints present relatively better impact toughness compared with AA7018, whether in the BM, FZ and FQZ. Besides, the
very similar trend is observed in the two welded joints that compared with FZ, the BM and FQZ have the largest and lowest Charpy
impact energy, respectively.
The considerable difference in notch sensitivity is closely related to microstructural features. As discussed in Section 3.2, the
severe solute segregation, intergranular network eutectics and microcracks result in the poor notch toughness of FQZ compared to
weld metal. Moreover, the eutectic products at the grain boundary within AA7018 FQZ are found to be continuous but rather
unconnected for RDE40 FQZ, as shown in Fig. 13. As a result, it is understandable that the impact toughness of FQZ within AA7018
welds is much worse with respect to RED40.

5. Fracture behavior

5.1. Static failure

It is argued that Al-Li alloy fusion welds are sensitive to the FQZ grain boundary cracking during welding and repairing [28]. The
static behavior is found to occur preferentially along the FQZ under both ambient and cryogenic temperature [20]. The mechanical
property degradation due to FQZ has been confirmed not only in Al-Li alloy butt welds but also in T joints [23]. It is found that the
FQZ behavior is closely related with the tensile strength and fracture characteristic of the joints [24]. Thus, the influence of FQZ on
the fracture behavior has gained increasing attention and interest in scientific research and engineering application. Fig. 24 presents
the failure behavior of laser welded 2060 Al-Li-Cu alloys under static tensile loading with different strain rates [13].
It is clear that the similar failure behavior can be observed in AA2060 laser welded joints under different strain rates. Specifically,
no typical relationship between the fracture position and strain rate is found. All the tested tensile specimens fail in the fusion zone.
The failure fracture is always initiated very close to FQZ near the upper joint and then the cracks propagate along approximately 45°

Table 1
The Charpy impact energy value of TIG welded RDE40 and AA7018 [21].
Energy absorbed, J·cm−2

Material BM FZ FQZ

RDE40 53.2 33.3 21.4


AA7018 47.3 28.3 10.0

60
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 24. The failure behavior of 2060 laser welded joints under different strain rates of: (a) 10−4/s and (b) 10−1/s [13].

with the tensile stress axis, eventually leading to a typical shear fracture mode in the weld center. The fracture path in Fig. 24 verifies
the previous statement proposed by Gutierrez and Lippold that FQZ plays a crucial role in both cracking susceptibility and structural
integrity [28].
With respect to Y mode weld under static loading, the plastic deformation initially occurs at the most serious softening zone and
here is the FQZ at the wider upper joint. It is known that FQZ is a narrow zone with high microstructure and hardness gradient,
necessarily leading to inhomogeneous plastic deformation [68]. With the increased plastic deformation accumulation, it is easy to
neck for the failure facture. The significant strength undermatching of the joints can result in strain localization in the softening
region, reducing the strength and ductility [13,43]. Fig. 25 presents the optical micrographs of laser welded AA2060 under the strain
level of 0.6%[13].
It is found that the cracks initiate approximately from FQZ despite of the existence of weld reinforcement and misalignment, and
then grow into the columnar crystal zone, as marked by the arrows. The optical micrograph of the weld cross section under the strain
level of 0.4% is also studied. However, no obvious surface cracks can be observed. The microscopic analysis implies that initial failure

Fig. 25. The optical micrographs of laser welded AA2060 under the strain level of 0.6%: (a) the weld cross section and (b) detail of the region
marked by a rectangle in (a) [13].

61
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 26. The fracture behavior of 2060-T8/2099-T83 double-sided laser welded T-joints under hoop tensile loading: (a) the optical micrograph of
failed weld cross section and (b) fracture characteristic of FQZ via the SEM [23].

may occur in the FQZ (the minimum hardness zone of the joints) as soon as the strain reaches about 0.6% during monotonic tensile
testing. It would be better if there is higher magnification characterization to accurately identify the cracking location and failure
behavior.
In addition to Al-Li alloy butt joints, the failure behavior along FQZ can also be observed in Al-Li alloy T-joints under static
loading such as hoop tensile test and compression testing. As shown in Fig. 26(a), it is found that the crack propagates from PMZ to
the local enrichment region of FQZ [23]. When correlating the fracture behavior with softening degree between PMZ and FQZ local
enrichment region, it is believed that the crack propagation path during hoop tensile loading is associated with the most severe
softening area near the lower fusion line.
Furthermore, SEM examination is performed to understand the fracture mode and characterize the fracture feature in detail, as
shown in Fig. 26(b). The smooth nondendritic equiaxed grain morphology and tiny Al2Cu intergranular particle are observed in the
FQZ, showing a typical intergranular fracture mechanism with secondary cracks. The formation of intergranular eutectic constituents
or particles as well as microcracks within FQZ can significantly induce the stress concentration, resulting in the local plastic de-
formation and then intergranular failure fracture.
Generally, the actual loading condition under the fuselage can be simulated via the compression test in the direction of stringer.
Fig. 27(a) shows the fracture behavior of 2060-T8/2099-T83 double-sided laser welded T-joints under compression testing [23]. It is
clear that the crack also propagates alternately in the FQZ, PMZ and their juncture area, which is quite similar to the failure behavior
during hoop tensile loading. The FQZ is characterized by the intergranular fracture mode and the transgranular fracture feature is
distinguished in the PMZ, as shown in Fig. 27(b). The intergranular fracture mode further confirms that the weak bonding strength is
generated at the grain boundaries of FQZ. Some large voids in the FQZ are most likely to be formed due to equiaxed grains pullout.

5.2. Fatigue damage

The narrow FQZ is well believed to contribute to cracking resistance reduction and mechanical property degradation under static
loading. While, the monotonic tensile or compression behavior is completely different from fatigue damage mechanism. The fatigue
failure always happens at the stress level much lower than the tensile strength, due to the cyclic loading nature applied to materials.

Fig. 27. The fracture behavior of 2060-T8/2099-T83 double-sided laser welded T-joints under compression testing: (a) the optical micrograph of
failed weld cross section and (b) fracture characteristic of FQZ and PMZ via the SEM [23].

62
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

The fatigue fracture is the most common source of engineering structure failure, accounting for more than 80–90% of all the me-
chanical failures [69]. The high-speed railway vehicles and aircrafts are usually subjected to dynamic loading in routine service. It is
essential to get a clear understanding about the FQZ effect on the fatigue performance including crack initiation and growth.
It is well-known that the fatigue cracking behavior is quite sensitive to the local stress/strain concentration. And those preferred
sites with high stress concentration generally facilitate the fatigue crack initiation, causing permanent damage upon cyclic loading
[70]. Subsequently, with the increased damage accumulation, the microscopic cracks propagate to form a dominant crack, which is
able to reach the critical size from where it grows unstably and then almost instantaneously leads to failure fracture. This is termed as
crack propagation stage, governed by fracture mechanics [71]. In general, the fatigue damage mechanism is very sensitive to mi-
crostructural features during the early and late stage of crack propagation [72].
The scientists usually explore the failure behavior by conventional surface-based characterization methods. Nevertheless, the
surface cracking behavior is not completely representative of the bulk material due to complex stress state and microstructural
feature. Furthermore, the surface-based or postmortem observation of fractured specimens in a destructive manner cannot precisely
identify the fatigue crack initiation and propagation. With the rapid development of third-generation synchrotron radiation X-ray
computed microtomography (SR-μCT), accurate detection and identification of various defects can be realized in terms of the di-
mension, morphology and location. And the spatial crack shape especially about the evolution process can also be nondestructively
obtained [73]. Combined with traditional experiments, a sound investigation on the interaction of microstructure, porosity and
cracking behavior is expected to reveal the fatigue performance of Al-Li alloy welded joints.
Ex situ fatigue testing when unloading for X-ray tomography certainly provides a nonconservative fatigue life assessment due to
the crack closure effect [74,75]. In order to realize in situ fatigue SR-μCT, a self-designed in situ tension/tension fatigue rig based on
the similar design principle as Buffière’s has been developed by Wu et al., which is able to be fully compatibly mounted at the
BL13W1 of Shanghai Synchrotron Radiation Facility (SSRF) and at the 4W1A of Beijing Synchrotron Radiation Facility (BSRF) [15].
The schematic diagram of in situ fatigue 3D X-ray imaging is presented in Fig. 28. The 3D visualization and characterization of fatigue
damage evolution in laser welded 2A97-T3 and hybrid laser welded 7020-T651 have been carried out via in situ fatigue SR-μCT.
For Al alloy fusion welded joints, gas pores inevitably generate due to the sharp reduction of supersaturated hydrogen during the
solidification [74]. The size, morphology and location of gas pores play a crucial role in the cracking behavior and fatigue perfor-
mance of welded components [76]. The presence of gas pores can reduce the effective load-carrying area and induce severe stress
concentration, generally acting as sites for crack initiation and then significantly degrading the fatigue performance [77]. The
porosity induced fatigue damage is quite common in Al alloy fusion welded structures.
It is found that fatigue cracks are easily initiated from an individual large-sized pore, multiple linked pores and clustered pores
near the weld surface (see Fig. 29) [15,73]. Generally, after grinding the surface, welding inherent defects like near surface pores are
able to expose to the environment [76]. In this case, the welding defects such as two linked pores with an equivalent diameter of
200 μm on the weld surface in Fig. 29(a) and multiple clustered pores on the surface in Fig. 29(b) can induce extremely serious stress
concentration. Upon cyclic loading, the local stress around the gas pores exceeds the yield strength of materials and localized plastic

Fig. 28. In situ SR-μCT with a fatigue testing machine in which the X-ray beam can pass through the polymer tube and specimen and permit an
arbitrary angle rotation. Total weight: 4.2 kg, dimension: 22 × 15 × 17 cm3, maximum frequency: 10 Hz, maximum load: 1.0kN.

63
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 29. The fatigue crack evolution in 7020-T651 hybrid laser welds via the high-resolution 3D SR-μCT: (a) the crack initiation from two linked
porosities on the weld surface and (b) the crack initiation from multiple clustered pores on the surface [15,73].

deformation happens, potentially inducing the short crack initiation. Once initiating, influenced by the microstructural feature and
porosity distribution, the crack propagation behavior presents discontinuity and inhomogeneity, primarily growing along the weld
center with dense pores.
As to hybrid laser welded 7020-T651 Al-Zn-Mg alloys with 0.13 wt% Zr, the width of FQZ is approximately 100 μm via the
metallographic observation, as shown in Fig. 3(a). It is known that FQZ is a narrow region with poor mechanical properties. However,
it is clear from Fig. 29 that the cracks initiate from gas pores near the weld surface instead of FQZ under fatigued loading and then
propagate along the weld center. It is concluded that FQZ close to the fusion line has little contribution to the cracking behavior in the
presence of obvious defects near the surface.
Fig. 30 presents that the fatigue cracks initiate from an undercut at the weld toe and then propagate approximately along FQZ
[15]. Here, we focus on the crack propagation behavior. On the one hand, the severe element segregation and resultant intergranular
continuous eutectic products and liquid films within FQZ give rise to stress concentration and also intergranular brittleness. And the
formation of microcracks and the absence of abundant precipitated phases provide very limited resistance to plastic deformation. On
the other hand, there exists a considerable deviation on the microstructure and hardness distribution at FQZ interfaces. The weak
zone with sharp microstructure and property gradient generally proves to be the crack propagation path [15,48]. Note that the
fatigue crack deviation and acceleration are not influenced by internal gas pores due to the small size and far away from cracks. Thus,
the crack discontinuity and inhomogeneity are mainly susceptible to microstructural features of FQZ.

64
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 30. The fatigue crack evolution inside 7020-T651 hybrid laser welded joints via the high-resolution SR-μCT at the BL13W1 of SSRF [15].

Although the gas pores are inevitably generated inside Al alloy fusion welds, while it is demonstrated that careful surface pre-
paration, suitable filler material and optimum welding parameter together with short pulse laser cleaning can effectively reduce the
gas porosity as low as possible, as shown in Fig. 31 [1,47,78,79]. For laser welded 2A97 Al-Li-Cu alloys, the fatigue crack initiates
from approximately FQZ in the absence of an individual large-sized pore or multiple linked pores near the surface, and then pro-
pagates along the FQZ. It is demonstrated that the heterogeneous micro-plastic deformation easily occurs due to the large micro-
structure gradient under cyclic loading, contributing to local strain concentration and crack initiation [80,81]. As a result, both sides
of FQZ with sharp microstructure and hardness gradient are considered as the major fatigue cracking sites.
Based on the above discussions, it is reasonably believed that there exists a competition mechanism between FQZ and internal
defects on the fatigue cracking behavior in the absence of obvious defects near the surface. More experimental investigations and
theoretical analyses are required to explore the competition effect, which is of great significance to understand the failure behavior of
FQZ and evaluate the fatigue property of Al-Li alloy welded joints. In addition, the effect of residual stress near FQZ on the fatigue
cracking cannot be neglected, which is also the prospective research of FQZ fatigue damage mechanism.
Due to the limitation to plastic zone size of crack tip and resistance to dislocation motion, grain boundaries are always considered
as an important factor to control microcrack propagation. It is generally believed that small-sized grains correspond to better fatigue
performance and fracture toughness [82]. However, some opposite conclusions are also proposed by some researchers [83]. The
contradiction is often caused by the fact that the grain boundary characteristic and crystal orientation at the crack tip are not fully
included in the experimental analysis. Moreover, the coarse intergranular precipitates, continuous eutectic products and liquid films
as well as the wide grain boundary precipitation-free zones have important effects on the mechanical property and fatigue strength,
giving the weak intergranular bonding force during the deformation and then intergranular fracture [84,85]. Accordingly, the small-
sized grains with more grain boundaries are believed to show the much worse fracture property.
Fig. 32 presents the fatigue crack propagation path in MIG welded Al-Zn-Mg-Cu alloy T-joints [48]. It is clear that the crack
propagates along the fusion line, mainly at the interface between fine equiaxed grains and coarse grain structures. Although there are

Fig. 31. The fatigue crack evolution in 2A97 laser welded joints via the high-resolution 3D SR-μCT at the BL13W1 of SSRF [47].

65
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Fig. 32. The fatigue crack evolution inside MIG welded Al-Zn-Mg-Cu alloy T-joints: (a) the failure path via the EBSD observation, (b) the fracture
morphology via the SEM observation and (c) detail of the region marked by a blue rectangle in (b) [48].

some corrosion products on the fracture surface, it is still found that the crack primarily propagates in the form of intergranular mode
and the whole failed surface is relatively flat. It is known that the nondendritic equiaxed grains are separated from partially melted
grains by the high angle grain boundaries (> 15°). The intergranular cracks tend to propagate along the high angle grain boundaries.
The weak intergranular bonding strength of FQZ leads to the poor resistance to cracking and failure fracture.
In addition to qualitative propagation behavior, the crack growth resistance of FQZ is investigated quantitively using compact
tension (CT) specimens. Fig. 33 illustrates the fatigue crack growth curves of FQZ in both RDE40 and AA7018 TIG welds [21]. The
backbone of fracture mechanics when descripting the long crack propagation is da/dN-△K diagram. No extended explanation is
needed here. The double-logarithmic diagram is consisted of three regions including the threshold region, the so-called Paris region
and the region where crack grows towards to fracture. Here we focus on the crack stable growth stage in Paris region.
It is found in Fig. 33 that RDE40 FQZ presents the lower crack growth rate da/dN for a given stress intensity factor range △K, i.e.
the slope of da/dN-△K curve is slightly lower compared with AA7018 FQZ. It may attribute to different microstructural features of
FQZ between the two TIG welded alloys, as discussed in Section 3.2. More experimental investigations and theoretical analyses are
required to further confirm and explain this point. Moreover, it is also necessary to compare the FQZ crack growth rate with other
zones of the joints like weld center and HAZ to get a clear and comprehensive understanding of the fatigue crack growth resistance of
FQZ.
The fatigue damage evolution of Al alloys mainly consists of the crack initiation, stage I crack growth and stage II crack growth.

Fig. 33. da/dN-△K curves of TIG welded AA7018 and RDE40 alloys [21].

66
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Concerning Al alloy welded joints, fatigue crack initiation easily occurs due to extremely high stress concentration around inherent
defects and microstructural imperfections [15,86]. The fatigue lifetime is thus primarily determined by the crack propagation instead
of initiation. The classical Paris equation is quite appropriate to describe the stage II crack growth rate rather than stage I. In order to
avoid overestimating the stage I growth rate, Zheng and Hirt put forward a Zheng-Hirt formula as below, which is able to predict the
crack growth rate during both stage I and stage II [87].
da
= B( K Kth )2
dN (3)
−2
where B is fatigue crack propagation coefficient and for Al alloys with fatigue striations, B = 15.9E . Here, the elastic modulus E of
FQZ is expected to accurately obtain via nano-indentation technology. The crack threshold ΔKth is considered as a critical point for
long crack propagation [88]. As is known that when ΔK is less than ΔKth, the crack growth rate is very slow or approximately zero.
Conversely, the crack grows rapidly and the crack propagation comes to the stage I and II. Yoder et al. found that the smaller grain
size corresponded to a lower ΔKth [89]. The intrinsic threshold ΔKth depends on material parameters including the grain size, yield
strength and stress ratio. Based on these influence factors and dislocation theory, Herold et al. proposed the following equation [90]:
Kth = 3.28(1 R) yd
1/2
(4)
where R is stress ratio, σy is yield strength and d is grain size. Thus, the relationship between the crack growth rate and microstructure
feature can be obtained by introducing Eq. (4) into (3), as is shown in Eq. (5).
da 1/2]2
= B[ K 3.28(1 R) yd
dN (5)
As is discussed that FQZ with extremely fine equiaxed grains presents the worse nanohardness and then yield strength. It is thus
concluded that the crack threshold ΔKth of FQZ is lower compared to other parts of the joints via Eq. (4). The lower ΔKth of FQZ
means that it is much easier for fatigue cracks to enter into the stage I and II long crack propagation. Based on Eq. (5), the comparison
of the whole crack growth rate and resistance within different regions of the joints can be quantitatively obtained by using the EBSD,
nano-indentation technology and crack growth rate testing in the future.

6. Calibration of FQZ

Since the first report of FQZ in the late 1990s, it has gained extensive attention in science and engineering. Based on foregoing
studies, it is concluded that the formation and presence of FQZ close to the fusion line have a very negative effect on the softening
behavior, corrosion property and failure fracture under both static and cyclic loadings. As a result, it is necessary to explore the
solutions to calibrate FQZ and reduce the adverse effect as low as possible.
Initially, researchers have tried to eliminate or minimize FQZ by optimizing welding process parameters. However, it is found that
over a wide range of heat input, achieved by varying travelling speed and welding current during autogenous TIG welding, it is not
possible to eliminate FQZ although the width and grain size vary considerably [28]. It is understandable because FQZ is confirmed in
many welding methods, such as the variable polarity plasma arc, TIG, LBW and EBW welds. Generally, the width of FQZ decreases
with the increase in heat input [5]. Note that simply adjusting welding parameters to minimize FQZ may consequently lead to other
welding problems [1,91,92].
According to the formation mechanism of FQZ, two potential solutions to calibrate FQZ are proposed. One is to adjust the
chemical composition in the base material and filler wire, mainly referring to Zr, Li, Sc and Er. The other is to enhance the molten
pool disturbance via the pulsed current, arc oscillation, current-assisted welding and high-gravity (centrifugal force) welding en-
vironment. Next, these calibration solutions related with FQZ will be introduced in detail.
It is well-known that Zr is the major element of FQZ formation. The FQZ is expected to be completely eliminated when removing
Zr from alloying compositions. However, Zr is beneficial to control the recrystallization and texture of wrought products and also the
minor addition of Zr is able to greatly improve the strength and ductility of materials [93]. Thus, it is not feasible to calibrate FQZ by
removing Zr from base metals. It is found that the FQZ width decreases with the reduction in Zr and Li of filler wires [9]. Never-
theless, Dev et al. propose that the width of FQZ is not linearly related to mechanical property degradation [21]. It means that further
modification needs to be performed on the microstructural features of FQZ.
The addition of rare earth metals such as Sc is well believed to effectively reduce the adverse effect of FQZ on the mechanical and
corrosion properties, and also contribute to considerable microstructure refinement of the welds. Dev et al. confirmed that the
addition of Sc (hypereutectic range 0.65 wt% Sc) into the commercial AA5556 filler could increase the solute solubility and decrease
the freezing range, drastically reducing the intergranular element segregation and precipitated phase. Combined with pulsed current
welding, FQZ was expected to be eliminated in RDE40 TIG welds [21,94–96].
Recently, Han et al. have found that FQZ can be suppressed inside 2060-T8/2099-T83 Al-Li alloy T-joints with a new type filler
CW3 (Al-6.2Cu-5.4Si) instead of conventional AA4047 (Al-12Si) [24]. It is mainly attribute to the modification of major intergranular
precipitate within FQZ, from T (AlLiSi) to T2 (Al6CuLi3). Based on the intergranular crack propagation behavior, it is believed that the
T2 phase presents better resistance to plastic deformation and intergranular bonding strength compared with T phase. Since the
failure fracture primarily occurs along the grain boundaries of FQZ under hoop tensile loading, both the strength and elongation of T-
joints with the filler CW3 are significantly improved compared to the filler AA4047.
The discussions above provide the FQZ calibration solutions by adjusting alloying compositions in the filler wire. The element

67
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

selection is not confined to Sc and Cu mentioned in the examples. According to the similar calibration principle, more elements can be
tried to further extend the calibration methods. Besides, the effect of molten pool disturbance on the formation and presence of FQZ is
discussed as below.
Firstly we focus on the pulsed current and arc oscillation welding technologies, as described in detail in Section 2
[21,25,44,45,97,98]. It is found that FQZ is able to be eliminated via the PC or AO welding, primarily due to the FQZ location
variation, from fusion line into the bulk pool. The enhanced molten pool convection together with the resultant reduction in tem-
perature gradient and peak temperature lead to the strong intermixing between nondendritic and dendritic grains as well as the
homogenization in alloying composition, generating less and discontinuous eutectic products at the grain boundaries. Thus, the
detrimental effect of FQZ on the mechanical and corrosion properties is reduced.
Aidun et al. investigated the effect of enhanced convection induced by a high-gravity welding environment on the FQZ in 2195-T8
welds [99]. The stationary (spot) bead-on-plate TIG welds were performed at 1, 5 and 10 G (1 G = 9.8 m·s−2) via a multi-gravity
research welding system (MGRWS). The high-gravity environment with enhanced convection is able to alter the thermal and fluid
flow conditions in the molten pool, forming a weld in which there is neither a stagnant boundary layer nor an unmixed zone
[100,101]. The heterogeneous nucleation particles are believed to be swept into the molten pool under the high-gravity condition
and completely dissolved. As a result, the formation condition of FQZ is destroyed. This method can effectively eliminate FQZ, but
whether it is feasible to apply into manufacturing welding components is not clear yet [11].
Recently, it is found that the FQZ width can be considerable reduced from 71.8 μm to 31.1 μm in current-assisted 2060-T8 laser
welds compared to conventional laser welds [102]. Here the direct current is introduced into the weld pool through the filler
AA5087. The introduction of direct current is expected to enhance the fluid flow as well as heat and mass transfer effect in the molten
pool, which is beneficial to accelerate the migration and dissolution of strengthening phases and thus reduce the FQZ width. More
welding experiments and microstructural evolution/mechanical behavior investigations are required to confirm the feasibility of
current-assisted welding on the FQZ calibration.

7. Conclusions

The interesting FQZ close to the fusion line is a common microstructural phenomenon inside fusion welded high-strength Al alloys
with Zr element, and it has been proved to play a crucial role in the service performance. Thus, the formation mechanism, micro-
structural nature, mechanical and fatigue behavior of FQZ are reviewed in this paper. The research proportion of these aspects and
some typical papers are listed in Fig. 34. Moreover, the types of Al-Li alloys and 7000 series Al alloys related to FQZ study as well as
references are summarized in Table 2.
It is clear from Fig. 34 that the comprehensive and systematic investigations have been performed on the formation mechanism of
FQZ. It is believed that the heterogeneous nucleation from Zr-containing precipitates, such as Al3Zr and Al3(Lix,Zr1−x) inside original
BM, should be the well-accepted formation mechanism of FQZ. The alloying composition, solidification substrate, welding parameter,
temperature as well as disturbance play an important role in the formation of FQZ. Note that a thorough understanding of the
formation mechanism is beneficial to analyze microstructural features and propose more reasonable and effective calibration solu-
tions.
The study concerning the microstructural features of FQZ is still not comprehensive, mainly referring to the microstructure,
element segregation and resultant eutectic product as well as solidification cracking. Thus, it is necessary to explore more micro-
structural characteristics such as the strain distribution via EBSD, element planar characterization via synchrotron radiation X-ray
fluorescence micro-scanning (SR-μXRF) and the distribution of strengthening phase and dislocation via TEM. Besides, it is worth
noting to compare the difference in microstructural features between FQZ and other parts of the joints to find the uniqueness of FQZ.
It is found that FQZ is the weakest part of the joints, giving the lowest hardness. As to the softening behavior of FQZ, it is believed
that the severe solute segregation and resultant relatively continuous eutectic products and very few strengthening phases lead to the
serious softening. Recently, a more reasonable local softening behavior of FQZ is proposed and it is mainly attributed to hot cracks
and a small amount of short strengthening phases within FQZ of T-joints. In order to further determine the softening mechanism, it is
recommended to adopt the combination of TEM and nano-indentation testing as well as atomic force microscope on the butt welds.

Fig. 34. The research topics and proportion related with FQZ as well as representative papers of each research field.

68
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

Table 2
The high-strength Al alloys related to the FQZ study and references.
Al-Li alloy welds 7000 series Al alloy welds

1441 Refs. [44,45] 7018 Ref. [21]


2060 Refs. [13,23,24,53,60,102]
2090 Ref. [9] 7020 Refs. [15,19,25]
2099 Refs. [23,24,53]
2195 Refs. [22,99,101] 7 N01 Refs. [17,36]
2196 Ref. [54]
2198 Ref. [54] RDE40 Ref. [21]
2A97 Refs. [5,10,47]
5A90 Refs. [33,41,43] Others Refs. [37–39,48]
8090 Refs. [25,27]
Weldalite049 Ref. [34]

The quasi-static tensile fracture preferentially proceeds along the FQZ inside Al-Li alloy butt welds and T-joints under both
ambient and cryogenic temperature, and the failure fracture presents intergranular fracture mode. The current research is limited to a
qualitative analysis of static failure behavior. I am very interested in the relationship between the joint strength, fracture type and
hardness value of FQZ, since the failure occurs along the FQZ. This may be a perspective to further understand the effect of FQZ on
the mechanical property of the joints.
Currently, still very few papers focus on the fatigue damage mechanism of FQZ inside Al alloy fusion welds. It may attribute to
that compared with FQZ, the welding defects such as an individual large-sized pore and multiple linked pores on the surface will
induce more serious stress concentration for cracking sites. Thus, there exists a competition mechanism between internal defects and
FQZ on the cracking behavior without obvious defects near the surface. And the competition effect is of great significance to un-
derstand the fatigue behavior of FQZ and evaluate the fatigue property of the joints. The actual microstructure-based finite element
modeling by combining synchrotron radiation X-ray imaging and numerical simulation is quite essential to quantitatively study the
competition behavior. In addition, the effect of residual stress near the FQZ cannot be neglected, which is also the prospective
research of FQZ fatigue damage mechanism.
Concerning the calibration solutions of FQZ, two feasible and effective research ideas are reported according to the formation
mechanism. One is to adjust the chemical composition in the base metal and filler wire and the other is to enhance the molten pool
disturbance. Based on the two research ideas, more specific and economic calibration solutions of FQZ can be put forward in the
future.
In summary, a clear and detailed understanding on the formation mechanism of FQZ has been established. However, the proper
evaluation of the effect of FQZ on mechanical and fatigue properties is not perfect yet. First, although the softening phenomenon of
FQZ has been well-confirmed, there is still no well-accepted softening mechanism. Then, the relationship between the joint strength,
fracture type and hardness value of FQZ is expected to better reveal the influence of FQZ on the failure behavior. Besides, to
quantitatively investigate the competition mechanism between internal defects and FQZ on the cracking behavior, the synchrotron
radiation X-ray imaging combined with finite element modelling is a feasible and effective method. All of the above studies will
ultimately serve as a basic reference to the engineering application. More reliable and economic calibration solutions of FQZ need to
be put forward.

Acknowledgements

Sincere thanks are given to the supports of the National Natural Science Foundation of China (11572267), the Sichuan Science
and Technology Program (2017JY0216), the Open Research Project of State Key Laboratory of Traction Power (2018TPL_T03), and
the Doctoral Innovation Fund Program of Southwest Jiaotong University.

Appendix A. Supplementary material

Supplementary data to this article can be found online at https://doi.org/10.1016/j.engfracmech.2019.01.013.

References

[1] Xiao RS, Zhang XY. Problems and issues in laser beam welding of aluminum-lithium alloys. J Manuf Process 2014;16(2):166–75.
[2] Dursun T, Soutis C. Recent developments in advanced aircraft aluminum alloys. Mater Des 2014;56(4):862–71.
[3] Gupta RK, Nayan N, Nagasireesha G, Sharma SC. Development and characterization of Al-Li alloys. Mater Sci Eng A 2006;420:228–34.
[4] Lavernia EJ, Srivatsan TS, Mohamed FA. Strength, deformation, fracture behavior and ductility of aluminum-lithium alloys. J Mater Sci 1990;25(2):1137–58.
[5] Ning J, Zhang LJ, Bai QL, Xin XQ, Niu J, Zhang JX. Comparison of the microstructure and mechanical performance of 2A97 Al-Li alloy joints between
autogenous and non-autogenous laser welding. Mater Des 2017;120:144–56.
[6] He EG, Liu J, Lee JY, Wang KH, Politis DJ, Chen L, et al. Effect of porosities on tensile properties of laser-welded Al-Li alloy: an experimental and modelling
study. Int J Adv Manuf Technol 2018;95:659–71.

69
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

[7] Ahn J, He EG, Chen L, Dear J, Davies C. The effect of Ar and He shielding gas on fiber laser weld shape and microstructure in AA 2024–T3. J Manuf Process
2017;29:62–73.
[8] Lavernia EJ, Grant NJ. Aluminum-lithium alloys. J Mater Sci 1987;22(5):1521–9.
[9] Lin DC, Wang GX, Srivatsan TS. A mechanism for the formation of equiaxed grains in welds of aluminum-lithium alloy 2090. Mater Sci Eng A 2003;351:304–9.
[10] Fu BL, Qin GL, Meng XM, Yang J, Yong Z, Zhen L. Microstructure and mechanical properties of newly developed aluminum-lithium alloy 2A97 welded by fiber
laser. Mater Sci Eng A 2014;617:1–11.
[11] Prasad NE, Gokhale AA, Wanhill RJH. Aluminum-lithium alloys: processing, properties and applications. Oxford: Butterworth-Heinemann; 2013.
[12] Balducci E, Ceschini L, Messieri S, Wenner S, Holmestad R. Effects of overaging on microstructure and tensile properties of the 2055 Al-Cu-Li-Ag alloy. Mater Sci
Eng A 2017;707:221–31.
[13] Tao W, Han B, Chen Y. Microstructural and mechanical characterization of aluminum-lithium alloy 2060 welded by fiber laser. J Laser Appl
2016;28(2):022409.
[14] Wu SC, Hu YN, Duan H, Yu C, Jiao HS. On the fatigue performance of laser hybrid welded high Zn 7000 alloys for next generation railway components. Int J
Fatigue 2016;91:1–10.
[15] Hu YN, Wu SC, Song Z, Fu YN, Yuan QX, Zhang LL. Effect of microstructural features on the failure behavior of hybrid laser welded AA7020. Fatigue Fract Eng
Mater 2018;41(9):2010–23.
[16] Zheng YC, Zhao ZH, Zhang Z, Zong WM, Dong C. Internal crack initiation characteristics and early growth behaviors for very-high-cycle fatigue of a titanium
alloy electron beam welded joints. Mater Sci Eng A 2017;706:311–8.
[17] Wang XM, Li B, Li MX, Huang C, Chen H. Study of local-zone microstructure, strength and fracture toughness of hybrid laser-metal-inert-gas-welded A7N01
aluminum alloy joint. Mater Sci Eng A 2017;688:114–22.
[18] Liu H, Shang DG, Liu JZ, Guo ZK. Fatigue life prediction based on crack closure for 6156 Al-alloy laser welded joints under variable amplitude loading. Int J
Fatigue 2015;72:11–8.
[19] Çam G, Koçak M. Microstructural and mechanical characterization of electron beam welded Al-alloy 7020. J Mater Sci 2007;42(17):7154–61.
[20] Gutierrez A, Lippold JC, Lin WG. Nondentritic equiaxed zone formation in aluminum-lithium welds. Mater Sci Forum 1996;217–222:1691–6.
[21] Dev S, Murty BS, Rao KP. Effects of base and filler chemistry and weld techniques on equiaxed zone formation in Al-Zn-Mg alloy welds. Sci Technol Weld
Joining 2008;13(7):598–606.
[22] Kostrivas A, Lippold JC. Fusion boundary microstructure evolution in aluminum alloys. Weld World 2006;50:24–34.
[23] Han B, Chen Y, Tao W, Lei ZL, Li H, Guo S, et al. Nano-indentation investigation on the local softening of equiaxed zone in 2060–T8/2099-T83 aluminum-
lithium alloys T-joints welded by double-sided laser beam welding. J Alloys Compd 2018;756:145–62.
[24] Han B, Tao W, Chen YB, Li H. Double-sided laser beam welded T-joints for aluminum-lithium alloy aircraft fuselage panels: Effects of filler elements on
microstructure and mechanical properties. Opt Laser Technol 2017;93:99–108.
[25] Reddy GM, Gokhale AA, Prasad KS, Rao KP. Chill zone formation in Al-Li alloy welds. Sci Technol Weld Joining 1998;3(4):208–12.
[26] Li XF, Li XH, Xiong HP, Li Y, Guo SQ, Zhang XJ. Microstructure and crack susceptibility of aluminum-lithium alloy weldment. J Aeronaut Mater
2007;27(3):55–9.
[27] Lee MF, Huang JC, Hou NJ. Microstructural and mechanical characterization of laser-beam welding of a 8090 Al-Li thin sheet. J Mater Sci 1996;31(6):1455–68.
[28] Gutierrez A, Lippold JC. A proposed mechanism for equiaxed grain formation along the fusion boundary in aluminum-copper-lithium alloys. Weld J
1998;77(3):123–32.
[29] Yang B, Xuan FZ, Chen ZK. Creep behavior of subzones in a CrMoV weldment characterized by the in-situ creep test with miniature specimens. Mater Sci Eng A
2018;736(24):193–201.
[30] Yang B, Xuan FZ, Liu XP. Heterogeneous creep behavior of a CrMoV multi-pass weld metal. Mater Sci Eng A 2017;690:6–15.
[31] Easterling KE. Solidification microstructure of fusion welds. Mat Sci Eng 1984;65(1):191–8.
[32] Davies GJ, Garland JG. Solidification structures and properties of fusion welds. Metall Rev 1975;20(1):83–108.
[33] Cui L, Li XY, He DY, Chen L, Gong SL. Effect of Nd: YAG laser welding on microstructure and hardness of an Al-Li based alloy. Mater Charact
2012;71(5):95–102.
[34] Shah SR, Wittig JE, Hahn GT. Microstructural analysis of a high strength Al-Cu-Li (Weldalite049) alloy weld. Proceedings of the third international conference
on trends in welding research. 1992. p. 281–5.
[35] Kostrivas A, Lippold JC. Weldability of Li-bearing aluminum alloys. Metall Rev 1999;44(6):217–37.
[36] Qiao JN, Lu JX, Wu SK. Fatigue cracking characteristics of fiber Laser-VPTIG hybrid butt welded 7N01P-T4 aluminum alloy. Int J Fatigue 2017;98:32–40.
[37] Yang DX, Li XY, He DY, Nie ZR, Huang H. Microstructural and mechanical property characterization of Er modified Al-Mg-Mn alloy Tungsten Inert Gas welds.
Mater Des 2012;34:655–9.
[38] Zhang L, Li XY, Nie ZR, Huang H, Sun JT. Microstructure and mechanical properties of a new Al-Zn-Mg-Cu alloy joints welded by laser beam. Mater Des
2015;83:451–8.
[39] Zhang L, Li XY, Nie ZR, Huang H, Sun JT, Sun ZG. Microstructure and Mechanical Properties of Joints of a New Al-Zn-Mg-Cu Alloy Welded by TIG. Rare Metal
Mat Eng 2016;45(3):698–701.
[40] Blake N, Hopkins MA. Constitution and age hardening of Al-Sc alloys. J Mater Sci 1985;20(8):2861–7.
[41] Cui L, Li XY, He DY, Chen L, Gong SL. Microstructure investigation of Nd:YAG laser welded 5A90 aluminium-lithium alloys. Trans China Weld Inst
2010;31(9):77–8.
[42] Rystad S, Ryum N. A metallographical investigation of the precipitation and recrystallization process in an Al-Zr alloy. Aluminum 1977;53(3):193–5.
[43] Xu F, Chen L, Gong SL, Li XY, Yang J. Microstructure and mechanical properties of Al-Li alloy by laser welding with filler wire. Rare Metal Mat Eng
2011;40(10):1775–9.
[44] Reddy GM, Gokhale AA, Rao KP. Effect of filler metal composition on weldability of Al-Li alloy 1441. Sci Technol Weld Joining 2014;3(3):151–8.
[45] Reddy GM, Gokhale AA, Rao KP. Effect of the ratio of peak and background current durations on the fusion zone microstructure of pulsed current gas tungsten
arc welded Al-Li alloy. J Mater Sci Lett 2002;21(20):1623–5.
[46] Wu SC, Hu YN, Song XP, Xue YL, Peng JF. On the Microstructural and mechanical characterization of hybrid laser-welded Al-Zn-Mg-Cu alloys. J Mater Eng
Perform 2015;24(4):1540–50.
[47] Chen L, Hu YN, He EG, Wu SC, Fu YN. Microstructural and failure mechanism of laser welded 2A97 Al-Li alloys via synchrotron tomography. Int J Lightweight
Mater Manuf 2018. https://doi.org/10.1016/j.ijlmm.2018.08.001.
[48] Liu C, Yuan DW, Yang XB, Liu GX, Qin F, Liu CH, et al. Effects of microstructural heterogeneity on crack behaviors in welding zones of aluminum alloy parts. J
Chin Electron Microscopy Soc 2015;34(3):181–8.
[49] Jian HG, Yin ZM, Xie XE, Jiang F. Microstructure and properties of welded joint of 5A01 aluminum alloy for marine. J Wuhan Univ Technol 2013;35(6):50–4.
[50] Tao W, Yang ZB, Shi CY, Dong DY. Simulating effects of welding speed on melt flow and porosity formation during double-sided laser beam welding of AA6056-
T4/AA6156-T6 aluminum alloy T-joint. J Alloys Compd 2017;699:638–47.
[51] Yang ZB, Tao W, Li LQ, Chen YB, Shi CY. Numerical simulation of heat transfer and fluid flow during double-sided laser beam welding of T-joints for aluminum
aircraft fuselage panels. Opt Laser Technol 2017;91:120–9.
[52] Enz J, Khomenko V, Riekehr S, Ventzke V, Huber N, Kashaev N. Single-sided laser beam welding of a dissimilar AA2024-AA7050 T-joint. Mater Des
2015;76:110–6.
[53] Han B, Tao W, Chen Y. New technique of skin embedded wire double-sided laser beam welding. Opt Laser Technol 2017;91:185–92.
[54] Enz J, Riekehr S, Ventzke V, Kashaev N. Influence of the local chemical composition on the mechanical properties of laser beam welded Al-Li alloys. Phys Proc
2012;39:51–8.
[55] Wu SC, Yu X, Zuo RZ, Zhang WH, Xie HL, Jiang JZ. Porosity, element loss, and strength model on softening behavior of hybrid laser arc welded Al-Zn-Mg-Cu

70
Y.N. Hu et al. Engineering Fracture Mechanics 208 (2019) 45–71

alloy with synchrotron radiation analysis. Weld J 2013;92(3):64–71.


[56] Dixit M, Mishra RS, Sankaran KK. Structure-property correlations in Al 7050 and Al 7055 high-strength aluminum alloys. Mater Sci Eng A 2008;478:163–72.
[57] Starink MJ, Wang SC. A model for the yield strength of overaged Al-Zn-Mg-Cu alloys. Acta Mater 2003;51(17):5131–50.
[58] Malarvizhi S, Balasubramanian V. Influences of welding processes and post-weld ageing treatment on mechanical and metallurgical properties of AA2219
aluminium alloy joints. Weld World 2012;56:105–19.
[59] Huang M, Li AG, Zhang K, Feng CH. Laser welding characteristics of 2A97 Al-Li alloy. Trans China Weld Inst 2014;35(5):100–4.
[60] Zhang XY, Huang T, Yang WX, Xiao RS, Liu Z, Li L. Microstructure and mechanical properties of laser beam-welded AA2060 Al-Li alloy. J Mater Process Technol
2016;237:301–8.
[61] Li H, Zou JS, Yao JS, Peng HP. The effect of TIG welding techniques on microstructure, properties and porosity of the welded joint of 2219 aluminum alloy. J
Alloys Compd 2017;727:531–9.
[62] Narayana GV, Sharma VMJ, Diwakar V, Kumar KS, Prasad RC. Fracture behavior of aluminum alloy 2219–T87 welded plates. Sci Technol Weld Joining
2004;9(2):121–30.
[63] Sun GQ, Niu JP, Chen YJ, Sun FY, Shang DG, Chen SJ. Experimental research on fatigue failure for 2219–T6 aluminum alloy friction stir-welded joints. J Mater
Eng Perform 2017;26(8):1–8.
[64] Hansen N. Hall-Petch relation and boundary strengthening. Scr Mater 2004;51:801–6.
[65] Heidarzadeh A, Saeid T. Correlation between process parameters, grain size and hardness of friction-stir-welded Cu-Zn alloys. Rare Met 2016:1–11.
[66] Shercliff HR, Ashby MF. A process model for age hardening of aluminium alloys-I. The model. Acta Metallurgica et Materialia 1990;38(10):1789–802.
[67] Chia TL, Easton MA, Zhu SM, Gibson MA, Birbilis N, Nie JF. The effect of alloy composition on the microstructure and tensile properties of binary Mg-rare earth
alloys. Intermetallics 2009;17(7):481–90.
[68] Wang RZ, Chen B, Zhang XC, Tu ST, Wang J, Zhang CC. The effects of inhomogeneous microstructure and loading waveform on creep-fatigue behaviour in a
forged and precipitation hardened nickel-based superalloy. Int J Fatigue 2017;97:190–201.
[69] Cheung MS, Li WC. Probabilistic fatigue and fracture analyses of steel bridges. Struct Saf 2003;25(3):245–62.
[70] Sangid MD. The physics of fatigue crack initiation. Int J Fatigue 2013;57(12):58–72.
[71] Pessard E, Bellett D, Morel F, Koutiri I. A mechanistic approach to the Kitagawa-Takahashi diagram using a multiaxial probabilistic framework. Eng Fract Mech
2013;109(3):89–104.
[72] Zhang C, Gao M, Zeng XY. Effect of microstructural characteristics on high cycle fatigue properties of laser-arc hybrid welded AA6082 aluminum alloy. J Mater
Process Technol 2016;231:479–87.
[73] Wu SC, Xiao TQ, Withers PJ. The imaging of failure in structural materials by synchrotron radiation X-ray microtomography. Eng Fract Mech 2017;182:127–56.
[74] Wu SC, Yu C, Zhang WH, Fu YN, Helfen L. Porosity induced fatigue damage of laser welded 7075–T6 joints investigated via synchrotron X-ray micro-
tomography. Sci Technol Weld Joining 2015;20(1):11–9.
[75] Zhu SP, Foletti S, Beretta S. Probabilistic framework for multiaxial LCF assessment under material variability. Int J Fatigue 2017;103:371–85.
[76] Wu SC, Yu C, Yu PS, Buffière JY, Helfen L, Fu YN. Corner fatigue cracking behavior of hybrid laser AA7020 welds by synchrotron X-ray computed micro-
tomography. Mater Sci Eng A 2016;651:604–14.
[77] Zerbst U, Ainsworth RA, Beier HT, Pisarski H, Zhang ZL, Nikbin K, et al. Review on fracture and crack propagation in weldments-A fracture mechanics
perspective. Eng Fract Mech 2014;132:200–76.
[78] Alshaer AW, Li L, Mistry A. The effects of short pulse laser surface cleaning on porosity formation and reduction in laser welding of aluminium alloy for
automotive component manufacture. Opt Laser Technol 2014;64:162–71.
[79] Ion JC. Laser beam welding of wrought aluminum alloys. Sci Technol Weld Joining 2000;5(5):265–76.
[80] Liu J, Gao XL, Zhang LJ, Zhang JX. A study of fatigue damage evolution on pulsed Nd: YAG Ti6Al4V laser welded joints. Eng Fract Mech 2014;117:84–93.
[81] Zhang WC, Zhu ML, Wang K, Xuan FZ. Failure mechanisms and design of dissimilar welds of 9%Cr and CrMoV steels up to very high cycle fatigue regime. Int J
Fatigue 2018;113:367–76.
[82] Zhai T, Wilkinson AJ, Martin JW. A crystallographic mechanism for fatigue crack propagation through grain boundaries. Acta Mater 2000;48(20):4917–27.
[83] Pao PS, Jones HN, Cheng SF, Feng CR. Fatigue crack propagation in ultrafine grained Al-Mg alloy. Int J Fatigue 2005;27(10):1164–9.
[84] Sha G, Wang YB, Liao XZ, Duan ZC, Ringer SP, Langdon TG. Microstructural evolution of Fe-rich particles in an Al-Zn-Mg-Cu alloy during equal-channel angular
pressing. Mater Sci Eng A 2010;527:4742–9.
[85] Chen YQ, Pan SP, Zhou MZ, Yi DQ, Xu DZ, Xu YF. Effects of inclusions, grain boundaries and grain orientations on the fatigue crack initiation and propagation
behavior of 2524–T3 Al alloy. Mater Sci Eng A 2013;580:150–8.
[86] Kutsuna M, Yan Q. Study on porosity formation in laser welds in aluminium alloys (Report 1): Effects of hydrogen and alloying elements. Weld Int
1998;12(12):937–49.
[87] Zheng XL, Hirt MA. Fatigue crack propagation in steels. Eng Fract Mech 1983;18(5):965–73.
[88] Zerbst U, Madia M, Vormwald M, Beier HT. Fatigue strength and fracture mechanics-a general perspective. Eng Fract Mec 2017;5:745–52.
[89] Yoder GR, Cooley LA, Crooker TW. Quantitative analysis of microstructural effects on fatigue crack growth in widmanstätten Ti-6A1-4V and Ti-8Al-1Mo-1V.
Eng Fract Mec 1979;11:805–16.
[90] Herold H, Streitenberger M, Zinke M, Orazi L, Cammarota GP. An experimental and theoretical approach for an estimation of ΔKth. Fatigue Fract Eng Mater
2010;23(9):805–12.
[91] Ascari A, Fortunato A, Orazi L, Campana G. The influence of process parameters on porosity formation in hybrid Laser-GMA welding of AA6082 aluminum
alloy. Opt Laser Technol 2012;44(5):1485–90.
[92] Casalino G, Mortello M, Leo P, Benyounis KY, Olabi AG. Study on arc and laser powers in the hybrid welding of AA5754 Al-alloy. Mater Des 2014;61(9):191–8.
[93] Rioja RJ, Liu J. The Evolution of Al-Li base products for aerospace and space applications. Metall Mater Trans A 2012;43(9):3325–37.
[94] Kramer LS, Tack WT, Fernandes MT. Scandium in aluminum alloys. Adv Mater Process 1997;152(4):23–4.
[95] Spierings AB, Dawson K, Heeling T, Uggowitzer PJ, Schäublin R, Palm F, et al. Microstructural features of Sc- and Zr-modified Al-Mg alloys processed by
selective laser melting. Mater Des 2017;115:52–63.
[96] Cross CE, Grong Ø, Mousavi M. A model for equiaxed grain formation along the weld metal fusion line. Scr Mater 1999;40(10):1139–44.
[97] Mohandas T, Reddy GM. A comparison of continuous and pulse current gas tungsten arc welds of an ultra high strength steel. J Mater Process Technol
1997;69(1):222–6.
[98] Qiu L, Yang CL, Lin SB, Fan CL. Effect of pulse current on microstructure and mechanical properties of variable polarity arc weld bead of 2219–T6 aluminum
alloy. Mater Sci Technol 2009;25(6):739–42.
[99] Aidun DK, Dean JP. Effect of enhanced convection on the microstructure of Al-Cu-Li welds. Weld J 1999;78(10):139–54.
[100] Kang N, Singh J, Kulkarni AK. Effects of gravitational orientation on the microstructural evolution of gas tungsten arc welds in an Al-4 wt% Cu alloy. J Mater Sci
2003;38(17):3579–89.
[101] Hou KH, Baeslack WA. Effect of solute segregation on the weld fusion zone microstructure in CO2, laser beam and gas tungsten arc welds in Al-Li-Cu alloy 2195.
J Mater Sci Lett 1996;15(3):208–13.
[102] An N, Zhang XY, Yang WX, Xiao RS. Electrical current assisted laser welding of 2060 aluminum-lithium alloy with filler wire. Trans China Weld Inst
2017;38(3):83–6.

71

You might also like