You are on page 1of 17

Additive Manufacturing 40 (2021) 101879

Contents lists available at ScienceDirect

Additive Manufacturing
journal homepage: www.elsevier.com/locate/addma

Research Paper

Process-structure-property relations for as-deposited solid-state additively


manufactured high-strength aluminum alloy
C.J.T. Mason a, R.I. Rodriguez b, D.Z. Avery c, B.J. Phillips c, B.P. Bernarding c, M.B. Williams c, S.
D. Cobbs b, J.B. Jordon c, P.G. Allison c, *
a
Pacific Northwest National Laboratory, WA 99352, USA
b
Metals and Ceramics R&D, Boeing Research and Technology, Huntsville, AL, USA
c
Department of Mechanical Engineering, The University of Alabama, Tuscaloosa, AL 35487, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Solid-state additive manufacturing methods provide innovative solutions to circumvent problems associated with
AFS-D additive manufacturing materials susceptible to hot cracking by avoiding liquid solid phase transformations. In this work, the process
Solid-state parameter influence on microstructural evolution and mechanical response of a fully dense aluminum alloy 7050
Aluminum alloy 7050
(AA7050) component manufactured via a rapid, solid-state additive manufacturing process known as Additive
EBSD
Tensile orientation
Friction Stir Deposition (AFS-D) was quantified for the first time. Three sections (starting dwell, transient,
crossover of roads) of the deposition that exhibit differing thermomechanical processing histories were evaluated
for the resulting microstructure and mechanical response. The microstructural characterization was performed
on the as-deposited AA7050 via Electron Backscatter Diffraction (EBSD), Transmission electron microscopy
(TEM), optical microscopy, and Scanning Electron Microscopy (SEM). The microstructural characterization
revealed refined constituent particles and grains throughout the as-deposited AA7050 microstructure. Further­
more, quasi-static tensile experiments were conducted in both the build and transverse directions, in order to
quantify the orientation influence on tensile properties of the as-deposited AA7050 build. Spatially dependent
tensile properties were observed in the material due to heat input variation coarsening of secondary phases
towards the initial layers of the AFS-D build. Post-mortem analysis revealed that voids nucleated and coalesced
from the overgrowth of the strengthening precipitates present in the material, resulting in fracture.

1. Introduction coalescence and growth around these cracked particles, which results in
fracture.
Aluminum Alloy 7050 (AA7050) is a precipitate hardened alloy that Fusion based and solid-state manufacturing techniques provide new
is used widely across aerospace and automotive industry due to exhib­ avenues to additively repair or manufacture complex aerospace com­
iting high strength to weight ratio, good machinability, and moderate ponents. To-date, limited research has been conducted on additively
corrosion resistance [1–8]. AA7050 is strengthened by matrix precipi­ manufactured 7XXX series alloys due to the materials predisposition to
tate structures composed of Guinier-Preston (GP) zones, η’ (MgZn2), and hot crack during the liquid-to-solid phase transformation that occurs in
η (MgZn2) phases. In addition to the strengthening precipitates, addi­ fusion based additive manufacturing. Although there was hot cracking
tional secondary phases, such as E-phase (Al18Mg3Cr2), S-phase present in the selective laser melting (SLM) manufactured AA7075
(Al2CuMg), and the T-phase (Al32(Mg,Zn)49) are found through the samples, Kaufman et al. was able to achieve 99% material density [15].
microstructure [9–11]. Silicon-based intermetallic (Mg2Si) and Iron rich Sistiaga et al. demonstrated that hot cracking can be reduced in SLM
(Al6(Fe,Mn), Al3Fe, Al(Fe,Mn,Si) and Al7Cu2Fe) particles are the main AA7075 when 4% wt. Si is added to the powder [16]. It was concluded
type of constituent particles found in wrought AA7050 [12,13]. by Reschetnik et al., that SLM AA7075 components cannot be used in
Wrought AA7050 consists of elongated, pancake shaped grain structures high performance application due to the material’s low mechanical
oriented with the rolling direction [14]. Under monotonic tensile properties and ductility [17]. Sinh et al. was able to produce fully dense
loading, constituent particles crack and nucleate voids followed by laser metal deposited AA7075 components without any defects.

* Corresponding author.
E-mail address: pallison@eng.ua.edu (P.G. Allison).

https://doi.org/10.1016/j.addma.2021.101879
Received 27 October 2020; Received in revised form 6 January 2021; Accepted 28 January 2021
Available online 1 February 2021
2214-8604/© 2021 Published by Elsevier B.V.
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 1. (A) First step (dwell section) of the AFS-D Process, where a rotating tool drives into the substrate heating up the material. (B) A layer is produced when the
tool moves across the substrate at a specified layer height while an actuator drives plastically deforming solid feedstock. (C) After a layer is completed, the spindle is
then raised and moves to a specified area to start the next section. This second layer plows through the previous layer to create the next deposition layer. (D) The
process is repeated until the structure reaches a desired build height. (E) Schematic overview illustrating each section that will be evaluated for microstructural and
mechanical properties variation. (F) Image of the completed AA7050 structure.

However, significant composition changes occurred due to the burnoff factors, continuous dynamic recrystallization and discontinuous
of Mg and Zn, in which Zn had about 62% burnoff and Mg had 44% [18]. recrystallization [31,34,39]. Gholami et al. investigated the micro­
A solid-state additive manufacturing technique, Additive Friction structural and mechanical response of FSP AA7075 and found that
Stir-Deposition (AFS-D), has the potential to circumvent the challenges FSPed AA7075 exhibited equiaxed grains 10–12 times smaller than that
produced in fusion based techniques. AFS-D is a nascent solid-state of the base material. This reduction in grain size resulted in an increase
technology to additively manufacture, repair, and join metals that de­ in hardness of about 31% [49]. Jata et al. performed an in depth study
rives from the hollow rotation tool friction stir cladding technique on mechanical property evolution in FSWed AA7050-T7451 [50]. They
developed and introduced by Stelt et al. and Liu et al. [19–21]. AFS-D found that the top of the weld nugget exhibited a higher hardness than
was first introduced by Rivera et al. [22] for Inconel alloys then later the bottom of the weld nugget. This was due to the temperature gradient
Phillips et al. [23] elucidated for the first time the influence of heat input during the welding process. The differences in temperature led to vari­
from varying process parameters on the resulting microstructure and ation in precipitate distribution from the top to the bottom of the weld
mechanical behavior for aluminum alloys. Specifically, AFS-D exhibits nugget. The temperatures exhibited in FSP and FSW control the refine­
similar fundamental physics to that of other solid phase processes like ment and distribution of precipitates in the weld nugget. Due to AA7XXX
equal-channel angular pressing (ECAP) [24,25], friction back extrusion series alloys being precipitate hardened, the precipitate size and distri­
[26], cold spray [27–30] or friction stir welding/processing bution correlate directly to a material’s mechanical properties [8].
(FSW)/(FSP) [31–33]. In particular, The AFS-D process is more similar Since AFS-D is a nascent technology, there has been limited research
to FSW and FSP then that of cold spray, in which recrystallized micro­ reported in the open literature on AFS-D manufactured materials [22,23,
structure and strong metallurgical bond is achieved through severe 44,45,51–53]. Rivera et al. first detailed the as-deposited properties and
plastic deformation caused by the hollow rotating tool [31,34–42]. One microstructure layer dependence for aluminum alloys by examining a 6
disadvantage of AFS-D compared to cold spray is that cold spray exhibits mm tall AA2219 build through the AFS-D process. The AA2219
a smaller step size allowing for thinner coatings or parts without component was fully dense and defect-free with equiaxed grain
excessive post machining [43]. However, one of the key advantages of morphology and a refined microstructure. AFS-D processed AA2219
the AFS-D process to cold spray is the ability for AFS-D to utilize solid exhibited a strong torsional fiber texture at the top of the build, and this
feed stock, metallic powder, or recycled machine chips to create an texture became weaker towards the middle and bottom sections of the
additive component [44]. In addition, AFS-D processed material exhibits build [51].
an equiaxed microstructure and good ductility in the build direction Phillips et al. investigated process-microstructure-property relations
[45]. of AA6061. Phillips et al. evaluated 4 mm tall AFS-D AA6061 builds with
The resulting weld nugget in FSW and FSP is similar to that of AFS-D 27 varying process parameters. The AFS-D AA6061 experienced a 20%
processed materials due to the dynamic recrystallization and grain reduction in Vickers hardness, for the optimized parameters, compared
refinement of the microstructure. This equiaxed grain microstructure in to the feedstock (AA6061-T651) material [23]. Similarly, Griffiths et al.
the weld nugget has increased mechanical properties compared to that [54] investigated the viability of the AFS-D process on repairing AA7075
of the base metal due to the Hall-Petch effect [31,34,46–48]. The dy­ substrate. However, the effect of process history on resulting micro­
namic recrystallization and grain refinement exhibit two main driving structural and mechanical properties was not investigated. More

2
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Table 1 of the build that underwent differing thermo-mechanical processing.


Qualitative heat input for each section of the deposition Henceforth, the three sections will be referred to as the starting dwell
[55]. section, transient section, and crossover section (Fig. 1F). Table 1 shows
Section location Heat input the qualitative heat input relative to each section of the deposition
Dwell section High
characterized based on past solid-state processing research [55]. The
Transient section Low dwell section where the tool shoulder imparts large frictional heat to
Crossover section Medium initially soften the material exhibits the highest heat input, followed by
the crossover section where the tool and deposition has to move through
a previously deposited road has the medium heat input, while the
recently, Avery et al. modeled the fatigue response of as-deposited
transient section contains the lowest heat input. Specifically, In the
AFS-D AA7075. The as-built AFS-D AA7075 exhibited a decrease in
dwell section where deposition begins, the hollow rotating tool begins
mechanical response compared to the initial feedstock. This was due to
each layer by plunging into the previous three layers or substrate
the coarsening of strengthening precipitates due to the AFS-D process
generating frictional heat to allow material flow to begin the deposition.
[53].
Once enough heat is generated, the rotating tool slowly raises up to the
The effect of thermal history induced by AFS-D process parameters
new layer height. After the desired height is reached, the rotating tool
on the resulting microstructure directly influences the mechanical
begins the specified tool path. As the tool traverses across the substrate
behavior and performance of AFS-D produced components. Due to
from left to right, it creates the transient section. Once the 178 mm long
AA7050 being a precipitate hardened alloy, the thermal history induced
path is completed, the rotating tool raises up while depositing material
by manufacturing of a large AFS-D component can affect the resulting
to fill in the exit hole creating a uniform layer height (1 mm), thus
material properties. To ascertain these effects, a systematic study was
resulting in completing the first layer indicated by Fig. 1B. To complete
performed on a solid-state additively manufactured AFS-D component in
Layer 2 (Fig. 1C), the same process is repeated. However, due to Layer 2
the through thickness direction of the component of three different
being at the same height as Layer 1, the AFS-D tool drives through the
sections of overlapping deposition regions containing different ther­
middle of Layer 1 removing the previously deposited layer, while
momechanical histories. To the best of the authors’ knowledge this was
simultaneously depositing material. This section is referred to as the
the first study on the microstructural and mechanical property depen­
crossover section. The AA7050 part was subjected to compressed air
dence on location and orientation of an as-deposited AA7050 structure.
cooling throughout the entire AFS-D process. After every other layer a
10-min stoppage in the AFS-D process was applied to allow the part to
2. Materials and methods
cool. The process was repeated layer by layer until the structure was
manufactured to be 38 mm tall.
Using a commercially available AFS-D B8 machine, AA7050-T7451
A process parameter study was performed prior to manufacturing the
9.5 × 9.5 mm square rod feedstock was deposited on 6.3 mm thick
AA7050 build, by depositing two 1 mm layers that are 88.9 mm long. As
AA7050-T7451 fixed substrate. The feedstock was fed through a 38.1
shown in Fig. 1A, builds were manufactured with feed rate (F) ranging
mm diameter, rotating tool that had four “teardrop” protrusions on the
from 50.8 to 94.0 mm/min, tool rotational speed (ω) varying from 220
face of the tool. These features generate frictional heat, induce high
to 360 RPM, and traversing velocity (V) ranging from 50.8 to
shear, and severe plastic deformation of the feedstock to metallurgically
177.8 mm/min.
bond the deposited material to the subsequent layers. Fig. 1 illustrates
For Vickers hardness measurements, the cross section of each build
how the AFS-D process produces fully dense layer-by-layer depositions.
section and initial feedstock was polished down to 3.0 µm diamond
A metallurgical bond between deposited layers is achieved by pushing
suspension. A microindentation grid with 0.45 × 0.45 mm indent
solid feedstock or metallic powder through a rotating, hollow tool. For
spacing on a Clemex CMT 0.2 kg indented the AFS-D cross-sections for
the present study, an AA7050 structure was additively manufactured
the dwell, transient, and crossover sections per ASTM E384-17 [56].
through the AFS-D process (Fig. 1E). There were three distinct sections
Metallographic samples were manufactured from the initial,

Fig. 2. Quasi-static specimen orientation in the cross-section of the build. (A) The top transverse orientation is taken close to the final layers of the build. The bottom
transverse orientation is taken towards the initial layers of the build. (B) The vertical orientation is taken from the build direction of the as-deposited
AA7050 structure.

3
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Table 2 intermediate and final layers for each build section. Standard grinding
Initial microstructure and hardness properties of the AA7050-T7451 elongated and polishing procedures down to 3 µm was carried out followed by
grain feedstock. vibratory polishing with 0.5 µm colloidal silica for 12 h. A Tescan Lyra
Specimen direction Average Vickers hardness (Hv) Average grain size (µm) FIB-FESEM coupled with an EDAX Octane Elite and Hikari Super EBSD
Longitudinal (L) 158 570
camera was utilized to perform energy–dispersive X-ray spectroscopy
Long transverse (LT) 151 28.0 (EDS) and electron backscatter diffraction (EBSD). The EDAX Hikari
Short transverse (ST) 155 49.0 Super EBSD camera performed the EDS and EBSD grain size analysis per
ASTM E2627-13 at a step size of 0.6 µm [57]. Subsequently, a FEI Tecnai
F-20 TEM imaged nanostructure features and the evolution of pre­
cipitates from the final layers to the initial layers of the transverse sec­
tion of the AFS-D AA7050 component. The resulting EBSD data was used
to generate pole figures (PF) and orientation distribution function
(ODF). Although XRD provides a more deterministic understanding of
the texture, there is sufficient confidence in the general character of the
EBSD generated PF and ODF. In order to evaluate the mechanical
response of the additively manufactured material, quasi-static tensile
samples were tested in the transverse and through thickness directions
using a subscale specimen geometry previously used for additive ma­
terial [58–60] (Fig. 2). Samples in the transverse direction were
machined from the top and bottom of the build to evaluate mechanical
response dependency on layer location in the deposition. Quasi-static
testing was performed using an MTS servo hydraulic load frame
equipped with a 20 kN load cell. Quasi-static tensile tests were per­
formed in triplicate at a constant strain rate of 0.001/s and the resulting
yield stress (YS) was evaluated via 0.2% offset. Strain was recorded
using a 5 mm extensometer. Upon fracture, quasi-static tensile speci­
mens underwent SEM fractography analysis.

Fig. 3. AFS-D process parameters evaluated for as-deposited AA7050 corre­


lating to deposition quality. The star is the parameters chosen for this study.

Fig. 4. (A) Locations in the build where hardness evaluation was performed with a corresponding Vickers hardness scale. (B) Crossover section hardness map. (C)
Traverse section hardness map. (D) Dwell section hardness map. The 0 in the X-position marks the center of the deposit.

4
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

shows the parameters that were evaluated as a function of weld pitch


and actuator feed rate. Weld pitch is crucial in identifying adequate
process parameters for AFS-D due to the transient nature of the process.
Each set of process parameters was evaluated for galling, voids, and
overfeeding similar to the process study performed by Phillips et al.
[23]. The selected parameters used to create the AA7050 structure were
a tool rotational speed of 225 RPM, 50.8 mm/min traversing speed, and
50.8 mm/min actuator speed, denoted by the star in Fig. 3.
Fig. 4 displays the hardness distribution of the AA7050 AFS-D build
(as-deposited) for the through thickness of the crossover section, tran­
sient section, and dwell section. All three sections exhibited a clear
degradation in hardness response, with the peak hardness located at the
final layers of the build that contained less heat input. The average
Vickers hardness for each section is shown in Fig. 5. The dwell section
exhibited the highest peak hardness of all three sections examined,
where the hardness in the dwell section ranges from ~135 HV to
~55 HV. It was observed that the band of high hardness starts softening
~8 mm from the final layers of the build. Additionally, at about 10 mm
down from the final layers in the dwell section there is a band of low
hardness that is about 5 mm thick. The transient section exhibits the
Fig. 5. Average Vickers hardness for each section. The error bars represent the lowest peak hardness of all three sections. The hardness in the transient
scatter associated with the maximum and minimum hardness values exhibited section ranges from ~115 HV to ~55 HV. The band of high hardness
in the deposited regions. starts softening ~5 mm from the top of the build. The crossover section
hardness ranges from ~133 HV to ~55 HV. The band of high hardness
3. Results and discussion starts softening ~8 mm from the final layers of the build, approximately
8 layers below the top surface of the deposit. The severe mechanical
3.1. Microstructure response gradient present within each section of the material is not
present in previous studies performed on as-deposited AFS-D AA6061
The initial microstructure of the AA7050-T7451 feedstock for AFS-D and as-deposited AFS-D AA2219 reported by Phillips et al. and Rivera
processing that was used in this present study is shown in Table 2. The et al., respectively [23,51]. The reduction in the hardness response
AA7050-T7451 feedstock exhibits larger grains in the longitudinal di­ exhibited in the as-deposited AFS-D AA7050 does not occur until
rection compared to the long transverse and short transverse direction. ~10 mm from the top of the build for all three sections. The absence of a
This is consistent with typical wrought microstructure exhibiting elon­ mechanical response gradient exhibited in prior studies on AFS-D
gated pancake shaped grains [5]. AA6061 and AFS-D AA2219 is due to the AFS-D components not
Prior to building the AFS-D AA7050 structure shown in Fig. 1F, a exceeding 10 mm in height. This suggested that once a component is
parametrization study was carried out. Twenty-seven varying parame­ additively fabricated to a height beyond 10 mm the repeated thermal
ters were evaluated, however, only one parameter was selected based on cycles begins to overage the material at the initial layers of the build.
quality of deposition (fully-dense deposition, no visible voids). Fig. 3 The mechanical response gradient exhibited in the as-deposited material

Fig. 6. Backscatter images of the precipitate distribution correlated to the relative location in the hardness plot for (A) final layers, (B) intermediate layers, and (C)
initial layers of the build crossover section.

5
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 7. (A) SEM image of the initial layers of the crossover section. (B) EDS analysis location of the precipitates from the crossover section initial layers. (C)
Magnesium distribution, (D) iron distribution, (E) copper distribution, (F) zinc distribution.

is consistent with the finding presented by Avery et al. [53]. strengthening precipitates and growth of secondary phases exhibited in
Backscatter images of the initial layers, intermediate layers, and final the initial layers is due to the increased thermal cycles from the AFS-D
layers of the crossover section are shown in Fig. 6. The images depicted process that the initial layers undergo compared to the final layers of
in Fig. 6 are representative of all three sections of the as-deposited AFS-D the as-built AFS-D AA7050 component.
AA7050 structure. A collection of intermetallic particles is located at the Fig. 10 shows EBSD images for the initial layers, intermediate layers,
grain boundaries throughout the build that is not exhibited in the and final layers of the crossover section, transient section, and dwell
feedstock. Additionally, the intermediate and initial layers of the section of the as-deposited AFS-D AA7050 component. The EBSD images
component contain a collection of intermetallic particles within the present an equiaxed grain structure throughout the build. Fig. 11 dis­
grains that are not present in the final layers of the as-built component. plays the cumulative grain size distribution for each section for the
EDS shows that the intermetallic particles along the grain boundaries initial, intermediate and final layers. The crossover section exhibited the
are Mg, Zn, and Cu rich (Fig. 7), whereas the intermetallic particles largest grains of all three sections. This increase in comparative grain
located within the grain are Mg and Zn rich. The particles present along size was attributed to the crossover section undergoing increased ther­
the grain boundaries are likely the S-phase due to the particles con­ mal cycles when compared to the transient and dwell sections of the
taining Cu and Mg. In contrast, the intermetallic particles present within component, which exhibit similar grain sizes. The transient and cross­
the grain are likely the overgrowth of η’ and η. The overgrowth of these over section displayed a slight decrease in grain size from the final layers
phases was due to the increase in thermal cycles that the initial and to the intermediate layers and then grain size increased towards the
intermediate layers of the build experienced compared to the final layers initial layers. However, in the dwell section, the grain size was smaller in
of the build. This hypotheses is supported by Fuller et al. where they the initial layers compared to the final layers. The reduction in grain size
performed TEM analysis on the heat affected zone of FSP AA7050 and at the initial layers compared to the final layers in the dwell section was
determined that there was overgrowth of η’ and η along the grain caused by relatively large force input coupled with the heat applied
boundaries and within the grain interior causing a decrease in me­ during the starting of a new layer. This large compressive force coupled
chanical response [48]. with severe plastic deformation of the grains leading to a reduction in
Scanning transmission electron microscopy-high angle annular dark grain size, a similar process to the Gleeble thermo-mechanical simulator
field (STEM-HAADF) of the TEM images from the final layers of the [65–68]. The final layers of the component exhibited similar grain size
component reveal that these intermetallic particles along the grain distribution for all three sections. However, towards the initial layers of
boundaries and within the grain are S-phase particles (Al2CuMg). There the build the disparity between grain size in each section grew due to the
are small T-phase (Al32(Mg,Zn)49) particles present within the grain as different thermal process histories that is subjected to each section.
well (Fig. 8). The needle like structures within the grain are η and η’ Furthermore, the dynamic recrystallized grains present throughout the
strengthening precipitates [61]. STEM-HAADF of TEM images from the component provides further evidence that the softening is caused by the
initial layers of the build reveal that there is significant growth of the over enlargement of the strengthening precipitates within the grain due
S-phase and T-phase particles within the grain interior (Fig. 9). The to overaging of the material from the thermal cycling of the AFS-D
precipitate size increased from 0.022 ± 0.002 µm to process.
0.06225 ± 0.012 µm from the final layers of the build to the initial Fig. 12 provides supplementary visualization of the dynamic
layers of the build, respectively (Table 3). There was also no evidence of recrystallization via grain reference orientation deviation (GROD) maps
η’ and η being present in the initial layers of the build. TEM images also of the initial, intermediate, and final layers for all three sections. The
revealed that there is an increase in precipitate free zones (PFZs) along GROD maps of AFS-D 6061 are similar to all three sections of the AFS-D
the grain boundaries from 0.32 ± 0.120 µm in the final layers of the AA7050 component [23]. The crossover exhibits the highest intra­
build to 0.509 ± 0.299 µm in the initial layers of the build (Table 3). granular misorientation followed by the transient section. High intra­
This overgrowth of phases and increases in PFZs along the grain granular misorientation is indicative of plastic deformation with
boundaries is the cause of the major reduction in mechanical response dislocation pileup. The crossover section experiences roughly the same
exhibited in the initial layers of the build [62–64]. The loss of intensity of intragranular misorientation from the final layers to the

6
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 8. (A) STEM-HAADF images from the final layers of the AFS-D AA7050 Component. (B) Mg overlay of particles. (C) Cu overlay of particles. (D) Zn overlay of
particles. (E) Fe overlay of particles. All scale bars are 0.5 µm in length unless otherwise specified.

initial layers of the build. However, the intensity of the intragranular Located in Table 3; Cube, Goss, Brass, and Copper are the four main FCC
misorientation for the intermediate layers of the transient and dwell rolling and recrystallization components. The Brass and Copper com­
decreases when compared to the final layers and then increases in value ponents are commonly exhibited in that of Rolled FFC material. The
when approaching the initial layers. Cube and Goss components result from recrystallization of the micro­
The common components of texture typically found in aluminum structure and is primarily exhibited in hot or cold rolled FCC material
alloys are shown in Table 3, where A, A*, C and B components are the [70].
ideal torsional texture found in aluminum alloys. In particular, A and A* Past studies on the resulting texture in FSW of aluminum alloys have
components are generated at low temperatures and strains. At large concluded that the texture in the stir zone is consistent with the ideal
strains, A and A* components are replaced by the C component. How­ torsion texture components [71–75]. In fact, the B and C torsional
ever, as temperature is raised under torsional loading at large strains the texture components are the primary texture components in the nugget
C component is then replaced by the A and A* components. Once tem­ zone of FSW aluminum alloys reported in literature. However, there are
peratures approach about 0.67 Tm at large strains the A and A* texture a few studies conducted on the resulting texture of FSW aluminum alloys
components are replaced by the B component. The B component results that exhibited Cube and Goss recrystallization texture components and
in a lower dislocation pile-up than that of the A and A* orientations [69]. did not contain B or C torsional texture components [76,77] (Table 4).

7
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 9. (A) STEM-HAADF images from the initial layers of the AFS-D AA7050 component. Red box denotes where EDS analysis was performed. (B) Mg overlay of
particles. (C) Cu overlay of particles. (D) Zn overlay of pareparticles. (E) Fe overlay of particles. All scale bars are 1 µm in length. (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this article.)

The ODF plots for the final layers of the crossover section demon­
Table 3 strate that the texture exhibited primarily contains the torsional texture
Average precipitate free zone length and average particle size from TEM images B component with a slight peak in the C component Fig. 13A. The final
of the initial and final layers of the build. layers of the crossover section texture response is similar to that of FSW
Location Average PFZ length (µm) Average precipitate size (µm) aluminum alloys [79]. The intermediate layers of the crossover section
Final layers 0.320 ± 0.120 0.022 ± 0.002 have the strongest texture of any location throughout the build. The ODF
Initial layers 0.509 ± 0.299 0.062 ± 0.012 plots for the intermediate layers in the crossover section exhibit very
sharp A and A* torsional components with a slight C component peak.
The initial layers of the crossover section contain a very sharp Cube
The EBSD generated PF and ODF shown in Fig. 13 display the vari­ component with a slight B torsional component texture. The texture in
ation in texture for the initial, intermediate, and final layers of the build the crossover section differs significantly from the top of the build to­
for all three sections. PF and ODF plots of the common texture compo­ wards the bottom. This stark difference in texture will result in
nents typically found in aluminum alloy are depicted in Fig. 14.

8
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 10. EBSD of the crossover section, transient section, and dwell section for the final layers, intermediate layers, and initial layers of the deposit. The scale bars are
35 µm in length.

anisotropic mechanical response in the crossover section of the as-built 3.2. Tensile behavior
AFS-D AA7050 structure.
The PF and ODF plots of the transient section for the final, inter­ Quasi-static (QS) tensile response in the orthogonal directions (top
mediate, and initial layers of the as-built AFS-D structure are located in transverse, bottom transverse, and vertical) for all three sections of the
Fig. 13B. The final layers of the transient section contain a moderate as-built AFS-D structure are shown in Fig. 15. A summary of the
texture response and include a combination of all four ideal torsional resulting strength coefficient, strain hardening exponent, YS, and ulti­
components. The intermediate layers of the transient section exhibit mate tensile strength (UTS) for the orthogonal directions for all three
very sharp B and A torsional texture components. In addition, there is sections are displayed in Fig. 16.
also a very slight Cube texture component response exhibited in the When compared to wrought AA7050-T7651 the as-built AA7050
intermediate layers. Lastly, the initial layers of the transient section yield strength for all orientations is significantly lower. However, the top
contain a very sharp Cube component response and do not exhibit any transverse orientation for all locations is roughly 50 MPa lower than that
components of the torsional texture components. Each layer deposited in of wrought AA7050-T7651. The reduction in mechanical response
the transient section contains a torsional texture. The previously compared to wrought values is due to the overgrowth of the strength­
deposited layers undergo a transition from a torsional texture to a Cube ening phases throughout the microstructure.
component recrystallization texture due to the repeated thermal cycles The top transverse sample orientation for the dwell section and the
and compressive forces applied from the AFS-D process. This difference crossover section exhibited similar YS and UTS response, while the
in texture could result in a variance in mechanical response between specimens located in the transient section experienced a 10.5% and
final and initial layers of the build. 7.4% reduction in YS and UTS, respectively. There was not a significant
The PF and ODF plots of the dwell section for the final, intermediate, difference in ductility for the top transverse orientation between each
and initial layers of the as-built AFS-D structure are located in Fig. 13C. section. The top transverse sample orientation exhibited the highest YS,
The final layers of the dwell section exhibit a moderate texture response. UTS, and elongation to failure when compared to the bottom transverse
The final layers contain slight B and C torsional texture components and and vertical sample orientation for each section. This increase in me­
a slight Cube texture response. The intermediate layers of the dwell chanical response coincides with the results from the microhardness
section exhibit a slight C torsional texture component and a slight Cube measurement as previously discussed.
texture component. The initial layers of the dwell section contain a For the bottom transverse sample orientation, the crossover section
strong B and A torsional component and a slight Cube and Goss exhibited the lowest YS and UTS when compared to the dwell and
component texture response. transient sections. All three sections demonstrated similar elongation to
failure. Additionally, the bottom transverse sample orientation was
slightly less ductile than that of the top transverse sample orientation.
The crossover section had the highest YS and UTS of all three sections

9
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 11. Cumulative grain size distribution size for the (A) crossover section, (B) transient section, and (C) dwell section for the initial layers, intermediate layers, and
final layers of the deposit. A comparison is given between each section for the (D) initial layers, (E) intermediate layers, and the (F) final layers.

in the vertical orientation. The transient section exhibited the lowest YS vertical sample orientation exhibited similar YS and UTS to the bottom
and UTS in the vertical sample orientation of all three sections. How­ sample orientation, the vertical sample orientation appeared to exhibit a
ever, the transient section was significantly more ductile in the vertical more brittle tensile response. The more brittle response is a byproduct of
sample orientation than the dwell and crossover section. Although the the mechanical response gradient observed in the hardness plots (Fig. 4).

10
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 12. Grain reference orientation deviation (GROD) plots of the cross sections of the initial, intermediate, and final layers for each of the three section in the
component. Misorientation angles greater than 10◦ are defined as high angle grain boundaries (HAGB) and are represented by thick black lines. The scale bars are
35 µm in length.

the final layers of the component was linked with the size increase of the
Table 4
η’ and η strengthening precipitates and secondary phases (Fig. 6). The
Components of texture commonly found in aluminum alloys [78–80].
increased size in precipitate phases within the grain function as stress
Texture orientations Indices Type concentrations leading to void nucleation and growth. Consequently,
A {11‾1‾} < 110 > Torsional due to dislocation pile-up, the amount of strain-energy present in the
{1‾11} < 1‾1‾0 > grains is diminished before the presence of localized void nucleation and
A* {1‾1‾1} < 112 > Torsional coalescence. The strain hardening for all three sections was approxi­
{111‾} < 112 >
C {001} < 110 > Torsional
mately twice that of FSP/FSW AA7050 reported in literature [14]. The
B {1‾12} < 110 > Torsional higher strain hardening exhibited in the AFS-D component relates to the
{11‾2‾} < 1‾1‾0 > lesser average dislocation density of recrystallized grains, shown in
Cube {100} < 001 > Recrystallization Fig. 12.
Goss {110} < 001 > Recrystallization
SEM images of the representative fractured surfaces for each tensile
Brass {110} < 11‾2 > Rolling
Copper {112} < 111‾ > Rolling sample orientation are depicted in Fig. 17, where Fig. 17A and B depict
the fracture surface for the vertical orientation from the crossover sec­
tion. Each orientation shows dimpling consistent with a ductile fracture.
During tensile loading, the bottom of the gauge section for the vertical The fracture surfaces for each sample orientation are similar despite
orientation specimen exhibited larger plastic deformation under each section and layer location exhibiting differing textures. This is
loading, while the top part of the gauge section only experienced slight likely due to the texture not being strong enough to alter the fracture
plastic deformation due to the bottom of the sample being weaker. As a paths. Fig. 17B shows that there is dimpling throughout the fracture
result, the non-uniform deformation occurring in the specimen provides surface and suggests that these dimples are the direct result of micro-
a reduction in elongation to failure in the vertical orientation. The void nucleation, followed by growth and coalescence from cracked
degradation in mechanical response for the initial layers compared to

11
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 13. EBSD generated PF and ODF plots (001 and 111 PF in stereographic projection and ϕ2 = 0◦ and ϕ2 = 45◦ ODF sections) of the as-deposited AFS-D AA7050
in the build direction representing the texture in the initial, intermediate, and final layers for the (A) crossover, (B) transient, and (C) dwell section. For the PFs the y
axis of the is the build direction (BD) and the x axis is the transverse direction (TD). The longitudinal direction is out of plane of the page.

12
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 14. Redrawn schematic representing common texture orientations for FCC aluminum as reported by Montheillet et al. [80] and Barlat [78] (A) PF representing
components A, A*, and C ideal torsional textures for FCC aluminum. (B) PF representing component B ideal torsional texture for FCC aluminum. (C) ODF at ϕ2
= 0◦ displaying the ideal torsional texture components locations. (D) ODF at ϕ2 = 45◦ displaying the ideal torsional texture components locations. (E) PF representing
the common rolling and recrystallization texture components for FCC aluminum. (F) ODF at ϕ2 = 45◦ displaying the common rolling and recrystallization texture
components location.

13
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

particles resulting in final fracture. In addition, the bottom transverse


orientation and top transverse orientation also displayed this fracture
mode on the fracture surface (Fig. 17C and D). The increase of size and
number of the constituent particles found within the as-deposited AFS-D
7050 material from the final layers to the initial layers of the deposition
was the driving force for the reduction in mechanical properties. These
large particles weaken the microstructure by causing large stress con­
centrations within the grains. This stress concentration nucleates and
coalesces voids causing fracture [81,82].

4. Conclusions

A fully dense AA7050 build with overlapping deposition paths


(178 mm × 178 mm x 38 mm) was successfully additively manufac­
tured through the rapid, thermomechanical, solid-state additive
manufacturing process, known as AFS-D. As such, the objective of this
research was to evaluate the microstructural and mechanical property
dependence on section location and build height for as-deposited AFS-D
AA7050. The dwell, transient, and crossover sections experienced
different thermo-mechanical histories during manufacturing of the AFS-
Fig. 15. Stress-strain response for as-deposited AFS-D AA7050 for the top
transverse, bottom transverse, and vertical specimen orientations. The error D AA7050 component resulting in observed differences in microstruc­
bars for true stress represent the max and min for each set of experiments. The ture and mechanical property. A summary of the main conclusions of
error bars for plastic strain is the range, which starts at the minimum value, for this study are presented below:
each tensile test.

Fig. 16. (A) Yield strength for each specimen orientation for the dwell, transient, and crossover sections. The red line is the reported wrought AA7050-T7651 value
[83]. (B) Ultimate tensile strength for each specimen orientation for the dwell, transient, and crossover sections. The red line is the reported wrought AA7050-T7651
value [83]. (C) Strength Coefficient for each specimen orientation for the dwell, transient, and crossover sections. (D) Strain hardening exponent for each specimen
orientation for the dwell, transient, and crossover sections. (For interpretation of the references to color in this figure legend, the reader is referred to the web version
of this article.) 14
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

Fig. 17. SEM images of the fractured quasi-static tensile specimens. (A) Fractured surface of the vertical tensile sample taken from the crossover section. (B)
Magnified SEM image of the ductile fracture surface from (A) denoted by the red square. (C) Fractured surface of the bottom transverse tensile sample taken from the
dwell section. (D) Magnified SEM image of the ductile fracture surface from (C) denoted by the red square. (E) Fractured surface of the top transverse tensile sample
taken from the transient section. (F) Magnified SEM image of the ductile fracture surface from (E) denoted by the red square. (For interpretation of the references to
color in this figure legend, the reader is referred to the web version of this article.)

• EDS analysis of TEM samples revealed that the initial layers exhibi­ thermal input occurring for each section in the build directly results
ted a significant growth of the T-phase and S-phase. Also, η and η’ in a texture variance exhibited throughout the structure.
were not found in the initial layers of the build. • The fractured surfaces of the specimens showed dimpling indicative
• The overgrowth of the strengthening precipitates present in the in­ of ductile fracture for all specimen orientations examined. Despite
termediate and initial layers for each section are the cause of mate­ the differing texture throughout the structure, the texture is too weak
rial softening due to overaging. The cause of the overaging is due to to influence the fracture paths during tensile loading.
the repeated thermal cycles that the initial and intermediate layers of • The resulting microstructure and mechanical properties of as-
the component underwent during to the AFS-D processing. deposited AA7050 is heavily dependent on thermo-mechanical pro­
• Despite the overaging microstructure exhibited in the initial layers, cessing history. This dependence must be accounted for when addi­
the as-deposited AFS-D AA7050 component exhibits dynamic tively manufacturing large components.
recrystallized equiaxed grains for all sections with GROD plots sug­ • Future studies need to be conducted on the solution treatment fol­
gesting relatively low dislocation pile-up would be present. lowed by aging of AFS-D AA7050 components in order to achieve
• The resulting texture in the as-built AFS-D AA7050 build varies peak mechanical properties in the material.
based on section location and layer location in the build. The varying

15
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

CRediT authorship contribution statement processability by selective laser melting, J. Mater. Process. Technol. 238 (2016)
437–445, https://doi.org/10.1016/j.jmatprotec.2016.08.003.
[17] W. Reschetnik, J.-P. Brüggemann, M.E. Aydinöz, O. Grydinc, K.-P. Hoyerc, Fatigue
C.J.T. Mason: Formal analysis, Writing - original draft, Investiga­ crack growth behavior and mechanical properties of EN AW-7075 aluminium
tion, Visualization. R.I. Rodriguez: Conceptualization, Supervision, alloy, in: 21st European Conference on Fracture, 2016, vol. 2, pp. 3040–3048, doi:
Writing - review & editing, Resources, Funding acquisition. D.Z. Avery: 10.1016/j.prostr.2016.06.380.
[18] A. Singh, A. Ramakrishnan, G. Dinda, Direct Laser Metal Deposition of Al 7050
Investigation, Writing - review & editing. B.J. Phillips: Investigation, Alloy, SAE Technical Paper, 2017, 2017-01-0286.
Writing - review & editing. B.P. Bernarding: Investigation, Formal [19] A.A. Van Der Stelt, T.C. Bor, H.J.M. Geijselaers, R. Akkerman, J. Huétink, Free
analysis. M.B. Williams: Investigation, Formal analysis. S.D. Cobbs: surface modeling of contacting solid metal flows employing the ALE formulation,
Key Eng. Mater. 504–506 (2012) 431–436, https://doi.org/10.4028/www.
Resources, Writing - review & editing. J.B. Jordon: Funding acquisition, scientific.net/KEM.504-506.431.
Project administration, Writing - review & editing, Conceptualization. P. [20] A.A. Van Der Stelt, T.C. Bor, H.J.M. Geijselaers, R. Akkerman, A.H. van den
G. Allison: Funding acquisition, Project administration, Writing - re­ Boogaard, Cladding of advanced Al alloys employing friction stir welding, Key Eng.
Mater. 554–557 (2013) 1014–1021.
view & editing, Conceptualization. [21] S. Liu, T.C. Bor, A.A. Van der Stelt, H. Geijselaers, C. Kwakernaak, A.M. Kooijman,
J. Mol, R. Akkerman, A.H. van den Boogaard, Friction surface cladding: an
Declaration of Competing Interest exploratory study of a new solid state cladding process, J. Mater. Process. Technol.
229 (2016) 769–784, https://doi.org/10.1016/j.jmatprotec.2015.10.029.
[22] O.G. Rivera, P.G. Allison, J.B. Jordon, O.L. Rodriguez, L.N. Brewer, Z. McClelland,
The authors declare that they have no known competing financial W.R. Whittington, D. Francis, J. Su, R.L. Martens, N. Hardwick, Microstructures
interests or personal relationships that could have appeared to influence and mechanical behavior of Inconel 625 fabricated by solid-state additive
manufacturing, Mater. Sci. Eng. A 694 (2017) 1–9, https://doi.org/10.1016/j.
the work reported in this paper. msea.2017.03.105.
[23] B.J. Phillips, D.Z. Avery, T. Liu, O.L. Rodriguez, C.J.T. Mason, J.B. Jordon, L.
Acknowledgements N. Brewer, P.G. Allison, Microstructure-deformation relationship of additive
friction stir-deposition Al–Mg–Si, Materialia 7 (2019), 100387, https://doi.org/
10.1016/j.mtla.2019.100387.
The authors would like to acknowledge Dr. Luke Brewer for his [24] B. Li, S. Joshi, K. Azevedo, E. Ma, K.T. Ramesh, R.B. Figueiredo, T.G. Langdon,
support and wisdom with texture analysis of the material. We gratefully Dynamic testing at high strain rates of an ultrafine-grained magnesium alloy
processed by ECAP, Mater. Sci. Eng. A 517 (1–2) (2009) 24–29, https://doi.org/
acknowledge the financial support of this research from the Boeing Co. 10.1016/j.msea.2009.03.032.
[25] Y.H. Zhao, X.Z. Liao, Z. Jin, R.Z. Valiev, Y.T. Zhu, Microstructures and mechanical
References properties of ultrafine grained 7075 Al alloy processed by ECAP and their
evolutions during annealing, Acta Mater. 52 (2004) 4589–4599, https://doi.org/
10.1016/j.actamat.2004.06.017.
[1] J.K. Park, A.J. Ardell, Microstructures of the commercial 7075 Al alloy in the T651
[26] M. Sarkari Khorrami, M. Movahedi, Microstructure evolutions and mechanical
and T7 tempers, Metall. Trans. A 14 (1983) 1957–1965, https://doi.org/10.1007/
properties of tubular aluminum produced by friction stir back extrusion, Mater.
BF02662363.
Des. 65 (2015) 74–79, https://doi.org/10.1016/j.matdes.2014.09.018.
[2] P.N. Adler, R. DeIasi, Calorimetric studies of 7000 series aluminum alloys: II.
[27] X.T. Luo, Y. Ge, Y. Xie, Y. Wei, R. Huang, N. Ma, C.S. Ramachandran, C.J. Li,
Comparison of 7075, 7050 and RX720 alloys, Metall. Trans. A 8A (1977)
Dynamic evolution of oxide scale on the surfaces of feed stock particles from
1185–1190, https://doi.org/10.1007/BF02667404.
cracking and segmenting to peel-off while cold spraying copper powder having a
[3] P.N. Adler, R. Deiasi, G. Geschwind, Influence of microstructure on the mechanical,
high oxygen content, J. Mater. Sci. Technol. 67 (2021) 105–115, https://doi.org/
Metall. Trans. 3 (1972) 3191–3200.
10.1016/j.jmst.2020.06.019.
[4] A.J. De Ardo, R.D. Townsend, The effect of microstructure on the stress-corrosion
[28] J.W. Murray, M.V. Zuccoli, T. Hussain, Heat treatment of cold-sprayed C355 Al for
susceptibility of a high purity AI-Zn-Mg alloy in a NaCI solution, Metall. Trans. 1
repair: microstructure and mechanical properties, J. Therm. Spray Technol. 27
(1970) 2573–2581.
(1–2) (2018) 159–168, https://doi.org/10.1007/s11666-017-0665-z.
[5] C.J.T. Mason, et al., Plasticity-damage modeling of strain rate and temperature
[29] Y.K. Wei, X.T. Luo, X. Chu, G.S. Huang, C.J. Li, Solid-state additive manufacturing
dependence of aluminum alloy 7075-T651, J. Dyn. Behav. Mater. 5 (1) (2019)
high performance aluminum alloy 6061 enabled by an in-situ micro-forging
868–874, https://doi.org/10.1007/s40870-019-00188-w.
assisted cold spray, Mater. Sci. Eng. A 776 (2020), 139024, https://doi.org/
[6] R.I. Rodriguez, J.B. Jordon, P.G. Allison, T. Rushing, L. Garcia, Microstructure and
10.1016/j.msea.2020.139024.
mechanical properties of dissimilar friction stir welding of 6061-to-7050 aluminum
[30] M.R. Rokni, C.A. Widener, G.A. Crawford, Microstructural evolution of 7075 Al gas
alloys, Mater. Des. 83 (2015) 60–65, https://doi.org/10.1016/j.
atomized powder and high-pressure cold sprayed deposition, Surf. Coat. Technol.
matdes.2015.05.074.
251 (2014) 254–263, https://doi.org/10.1016/j.surfcoat.2014.04.035.
[7] J. Zhang, Y. Deng, W. Yang, S. Hu, X. Zhang, Design of the multi-stage quenching
[31] J.-Q. Su, T.W. Nelson, C.J. Sterling, Microstructure evolution during FSW/FSP of
process for 7050 aluminum alloy, Mater. Des. 56 (2014) 334–344, https://doi.org/
high strength aluminum alloys, Mater. Sci. Eng. A 405 (1–2) (2005) 277–286,
10.1016/j.matdes.2013.09.029.
https://doi.org/10.1016/j.msea.2005.06.009.
[8] Y.L. Wang, Q.L. Pan, L.L. Wei, B. Li, Y. Wang, Effect of retrogression and reaging
[32] J.Q. Su, T.W. Nelson, R. Mishra, M. Mahoney, Microstructural investigation of
treatment on the microstructure and fatigue crack growth behavior of 7050
friction stir welded 7050-T651 aluminium, Acta Mater. 51 (3) (2003) 713–729,
aluminum alloy thick plate, Mater. Des. 55 (2014) 857–863, https://doi.org/
https://doi.org/10.1016/S1359-6454(02)00449-4.
10.1016/j.matdes.2013.09.063.
[33] R.S. Mishra, M.W. Mahoney, S.X. Mcfadden, N.A. Mara, A.K. Mukherjee, High
[9] A. Abolhasani, A. Zarei-Hanzaki, H.R. Abedi, M.R. Rokni, The room temperature
strain rate superplasticity in a friction stir processed 7075 Al alloy.
mechanical properties of hot rolled 7075 aluminum alloy, Mater. Des. 34 (2012)
[34] J.-Q. Su, T.W. Nelson, C.J. Sterling, Grain refinement of aluminum alloys by
631–636, https://doi.org/10.1016/J.MATDES.2011.05.019.
friction stir processing, Philos. Mag. 86 (1) (2006) 1–24, https://doi.org/10.1080/
[10] M.R. Rokni, A. Zarei-Hanzaki, A.A. Roostaei, A. Abolhasani, Constitutive base
14786430500267745.
analysis of a 7075 aluminum alloy during hot compression testing, Mater. Des. 32
[35] J.J. Pang, F.C. Liu, J. Liu, M.J. Tan, D.J. Blackwood, Friction stir processing of
(2011) 4955–4960, https://doi.org/10.1016/j.matdes.2011.05.040.
aluminium alloy AA7075: microstructure, surface chemistry and corrosion
[11] J. Albrecht, A.W. Thompson, I.M. Bernstein, The role of microstructure in
resistance, Corros. Sci. 106 (2015) 217–228, https://doi.org/10.1016/j.
hydrogen-assisted fracture of 7075 aluminum, Metall. Trans. A 10 (11) (1979)
corsci.2016.02.006.
1759–1766, https://doi.org/10.1007/BF02811712.
[36] M. Navaser, M. Atapour, Effect of friction stir processing on pitting corrosion and
[12] K.N. Solanki, M.F. Horstemeyer, W.G. Steele, Y. Hammi, J.B. Jordon, Calibration,
intergranular attack of 7075 aluminum alloy, 2017, doi: 10.1016/j.jmst.2016
validation, and verification including uncertainty of a physically motivated
.07.008.
internal state variable plasticity and damage model, Int. J. Solids Struct. 47 (2)
[37] A.H. Feng, D.L. Chen, Z.Y. Ma, Microstructure and cyclic deformation behavior of a
(2010) 186–203, https://doi.org/10.1016/j.ijsolstr.2009.09.025.
friction-stir-welded 7075 Al alloy, Metall. Mater. Trans. A 41 (4) (2010) 957–971,
[13] H.E. Hu, L. Zhen, L. Yang, W.Z. Shao, B.Y. Zhang, Deformation behavior and
https://doi.org/10.1007/s11661-009-0152-3.
microstructure evolution of 7050 aluminum alloy during high temperature
[38] M.W. Mahoney, C.G. Rhodes, J.G. Flintoff, W.H. Bingel, R.A. Spurling, Properties
deformation, Mater. Sci. Eng. A 488 (2008) 64–71, https://doi.org/10.1016/j.
of friction-stir-welded 7075 T651 aluminum, Metall. Mater. Trans. A 29 (1998)
msea.2007.10.051.
1955–1964, https://doi.org/10.1007/s11661-998-0021-5.
[14] B.C. White, R.I. Rodriguez, A. Cisko, J.B. Jordon, P.G. Allison, T. Rushing,
[39] S. Palanivel, P. Nelaturu, B. Glass, R.S. Mishra, Friction stir additive manufacturing
L. Garcia, Effect of heat exposure on the fatigue properties of AA7050 friction stir
for high structural performance through microstructural control in an Mg based
welds, J. Mater. Eng. Perform. 27 (2018) 3007–3013, https://doi.org/10.1007/
WE43 alloy, Mater. Des. 65 (2015) 934–952, https://doi.org/10.1016/j.
s11665-018-3379-6.
matdes.2014.09.082.
[15] N. Kaufmann, M. Imran, T.M. Wischeropp, C. Emmelmann, S. Siddique, F. Walther,
[40] M. Yuqing, K. Liming, H. Chunping, L. Fencheng, L. Qiang, Formation
Influence of process parameters on the quality of aluminium alloy en AW 7075
characteristic, microstructure, and mechanical performances of aluminum-based
using Selective Laser Melting (SLM), 2016, doi: 10.1016/j.phpro.2016.08.096.
components by friction stir additive manufacturing, Int. J. Adv. Manuf. Technol. 83
[16] M.L. Montero Sistiaga, R. Mertens, B. Vrancken, X. Wang, B. Van Hooreweder, J.
(9–12) (2016) 1637–1647, https://doi.org/10.1007/s00170-015-7695-9.
P. Kruth, J. Van Humbeeck, Changing the alloy composition of Al7075 for better

16
C.J.T. Mason et al. Additive Manufacturing 40 (2021) 101879

[41] M. Shahzad, M. Chaussumier, R. Chieragatti, C. Mabru, F. Rezai-Aria, Surface Mater. Sci. Eng. A 527 (2010) 2233–2240, https://doi.org/10.1016/j.
characterization and influence of anodizing process on fatigue life of Al 7050 alloy, msea.2009.11.057.
Mater. Des. 32 (6) (2011) 3328–3335, https://doi.org/10.1016/j. [63] J.D. Robson, Microstructural evolution in aluminium alloy 7050 during processing,
matdes.2011.02.027. Mater. Sci. Eng. A 382 (1–2) (2004) 112–121, https://doi.org/10.1016/j.
[42] F. Rui-dong, S. Zeng-qiang, S. Rui-cheng, L. Ying, L. Hui-jie, L. Lei, Improvement of msea.2004.05.006.
weld temperature distribution and mechanical properties of 7050 aluminum alloy [64] R. Ghiaasiaan, B.S. Amirkhiz, S. Shankar, Quantitative metallography of
butt joints by submerged friction stir welding, Mater. Des. 32 (10) (2011) precipitating and secondary phases after strengthening treatment of net shaped
4825–4831, https://doi.org/10.1016/j.matdes.2011.06.021. casting of Al-Zn-Mg-Cu (7000) alloys, Mater. Sci. Eng. A 698 (2017) 206–217,
[43] B.C. White, W.A. Story, L.N. Brewer, J.B. Jordon, Fatigue behavior of freestanding https://doi.org/10.1016/j.msea.2017.05.047.
AA2024 and AA7075 cold spray deposits, Int. J. Fatigue 112 (2018) 355–360, [65] Q. Guo, D. Li, H. Peng, S. Guo, J. Hu, P. Du, Nucleation mechanisms of dynamic
https://doi.org/10.1016/J.IJFATIGUE.2018.03.007. recrystallization in Inconel 625 superalloy deformed with different strain rates,
[44] J.B. Jordon, P.G. Allison, B.J. Phillips, D.Z. Avery, R.P. Kinser, L.N. Brewer, C. Cox, Rare Met. 31 (3) (2012) 215–220, https://doi.org/10.1007/s12598-012-0494-7.
K. Doherty, Direct recycling of machine chips through a novel solid-state additive [66] Y. Yang, Z.P. Xie, Z. Zhang, X. Li, Q. Wang, Y. Zhang, Processing maps for hot
manufacturing process, Mater. Des. 193 (2020), 108850, https://doi.org/10.1016/ deformation of the extruded 7075 aluminum alloy bar: anisotropy of hot
j.matdes.2020.108850. workability, Mater. Sci. Eng. A 615 (2014) 183–190, https://doi.org/10.1016/j.
[45] B.A. Rutherford, et al., Effect of thermomechanical processing on fatigue behavior msea.2014.07.072.
in solid-state additive manufacturing of Al-Mg-Si alloy, Metals 10 (7) (2020) 1–17, [67] G.Z. Quan, K.W. Liu, J. Zhou, B. Chen, Dynamic softening behaviors of 7075
https://doi.org/10.3390/met10070947. aluminum alloy, Trans. Nonferr. Met. Soc. China 19 (2009) s537–s541, https://doi.
[46] A. Goloborodko, T. Ito, X. Yun, Y. Motohashi, G. Itoh, Friction stir welding of a org/10.1016/S1003-6326(10)60104-5.
commercial 7075-T6 aluminum alloy: grain refinement, thermal stability and [68] D. Li, Q. Guo, S. Guo, H. Peng, Z. Wu, The microstructure evolution and nucleation
tensile properties, Mater. Trans. 45 (2004) 2503–2508. mechanisms of dynamic recrystallization in hot-deformed Inconel 625 superalloy,
[47] A. Gerlich, G. Avramovic-Cingara, T.H. North, Stir zone microstructure and strain Mater. Des. 32 (2) (2011) 696–705, https://doi.org/10.1016/j.
rate during Al 7075-T6 friction stir spot welding, Metall. Mater. Trans. A 37 (9) matdes.2010.07.040.
(2006) 2773–2786, https://doi.org/10.1007/BF02586110. [69] F. Montheillet, P. Gilormini, J.J. Jonas, Relation between axial stresses and texture
[48] C.B. Fuller, M.W. Mahoney, M. Calabrese, L. Micona, Evolution of microstructure development during torsion testing: a simplified theory, Acta Metall. 33 (4) (1985)
and mechanical properties in naturally aged 7050 and 7075 Al friction stir welds, 705–717, https://doi.org/10.1016/0001-6160(85)90035-5.
Mater. Sci. Eng. A 527 (9) (2010) 2233–2240, https://doi.org/10.1016/j. [70] Y. Zhou, J.J. Jonas, J. Savoie, A. Makinde, S.R. MacEwen, Effect of texture on
msea.2009.11.057. earing in FCC metals: finite element simulations, Int. J. Plast. 14 (1–3) (1998)
[49] S. Gholami, E. Emadoddin, M. Tajally, E. Borhani, Friction stir processing of 7075 117–138, https://doi.org/10.1016/S0749-6419(97)00044-2.
Al alloy and subsequent aging treatment, Trans. Nonferr. Met. Soc. China Engl. Ed. [71] B. Wang, B.B. Lei, J.X. Zhu, Q. Feng, L. Wang, D. Wu, EBSD study on
25 (9) (2015) 2847–2855, https://doi.org/10.1016/S1003-6326(15)63910-3. microstructure and texture of friction stir welded AA5052-O and AA6061-T6
[50] K.V. Jata, K.K. Sankaran, J.J. Ruschau, Friction-stir welding effects on dissimilar joint, Mater. Des. 87 (2015) 593–599, https://doi.org/10.1016/j.
microstructure and fatigue of aluminum alloy 7050-T7451, Metall. Mater. Trans. A matdes.2015.08.060.
31A (2000). Accessed 01 October 2018, 〈https://link.springer.com/content/pdf/1 [72] C. Zhang, G. Huang, Y. Cao, Y. Zhu, Q. Liu, On the microstructure and mechanical
0.1007%2Fs11661-000-0136-9.pdf〉. properties of similar and dissimilar AA7075 and AA2024 friction stir welding
[51] O.G. Rivera, P.G. Allison, L.N. Brewer, O.L. Rodriguez, J.B. Jordon, T. Liu, W. joints: effect of rotational speed, J. Manuf. Process. 37 (2019) 470–487, https://
R. Whittington, R.L. Martens, Z. McClelland, C.J.T. Mason, L. Garcia, J.Q. Su, doi.org/10.1016/j.jmapro.2018.12.014.
N. Hardwick, Influence of texture and grain refinement on the mechanical behavior [73] D.P. Field, T.W. Nelson, Y. Hovanski, K.V. Jata, Heterogeneity of crystallographic
of AA2219 fabricated by high shear solid state material deposition, Mater. Sci. Eng. texture in friction stir welds of aluminum, Metall. Mater. Trans. A Phys. Metall.
A 724 (2018) 547–558, https://doi.org/10.1016/j.msea.2018.03.088. Mater. Sci. 32 (11) (2001) 2869–2877, https://doi.org/10.1007/s11661-001-
[52] D.Z. Avery, O.G. Rivera, C.J.T. Mason, B.J. Phillips, J.B. Jordon, J. Su, 1037-2.
N. Hardwick, P.G. Allison, Fatigue behavior of solid-state additive manufactured [74] U.F.H.R. Suhuddin, S. Mironov, Y.S. Sato, H. Kokawa, Grain structure and texture
inconel 625, JOM 70 (2018) 2475–2484, https://doi.org/10.1007/s11837-018- evolution during friction stir welding of thin 6016 aluminum alloy sheets, Mater.
3114-7. Sci. Eng. A 527 (7–8) (2010) 1962–1969, https://doi.org/10.1016/j.
[53] D.Z. Avery, B.J. Phillips, C.J.T. Mason, M. Palermo, M.B. Williams, C. Cleek, O. msea.2009.11.029.
L. Rodriguez, P.G. Allison, J.B. Jordon, Influence of grain refinement and [75] W. Xu, J. Liu, G. Luan, C. Dong, Temperature evolution, microstructure and
microstructure on fatigue behavior for solid-state additively manufactured Al-Zn- mechanical properties of friction stir welded thick 2219-O aluminum alloy joints,
Mg-Cu alloy, Metall. Mater. Trans. A 51 (2020) 2778–2795, https://doi.org/ Mater. Des. 30 (6) (2009) 1886–1893, https://doi.org/10.1016/j.
10.1007/s11661-020-05746-9. matdes.2008.09.021.
[54] R.J. Griffiths, D.T. Petersen, D. Garcia, H.Z. Yu, Additive friction stir-enabled solid- [76] S. Chen, X. Jiang, Texture evolution and deformation mechanism in friction stir
state additive manufacturing for the repair of 7075 aluminum alloy, Appl. Sci. 9 welding of 2219Al, Mater. Sci. Eng. A 612 (2014) 267–277, https://doi.org/
(17) (2019) 3486, https://doi.org/10.3390/app9173486. 10.1016/j.msea.2014.06.014.
[55] R. Mishra, M.W. Mahoney, Y. Sato, Y. Hovanski, R. Verma, Friction stir welding [77] S.S. Yutaka, H. Hiroyuki, K. Ikeda, M. Enomoto, S. Jogan, T. Hashimoto,
and processing VII, 2016. Microtexture in the friction-stir weld of an aluminum alloy, Metall. Mater. Trans. A
[56] ASTM, E384-17 Standard Test Method for Microindentation Hardness of Materials, Phys. Metall. Mater. Sci. 32 (4) (2001) 941–948, https://doi.org/10.1007/s11661-
ASTM International, pp. 1–40, 2017, doi: 10.1520/E0384-17. 001-0351-z.
[57] ASTM, Standard Practice for Determining Average Grain Size Using Electron [78] F. Barlat, O. Richmond, Prediction of tricomponent plane stress yield surfaces and
Backscatter Diffraction (EBSD) in Fully Recrystallized Polycrystalline Materials, associated flow and failure behavior of strongly textured f.c.c. polycrystalline
ASTM International, pp. 1–5, 2014, doi: 10.1520/E2627. sheets, Mater. Sci. Eng. 95 (C) (1987) 15–29, https://doi.org/10.1016/0025-5416
[58] P.G. Allison, H. Grewal, Y. Hammi, H.R. Brown, W.R. Whittington, M. (87)90494-0.
F. Horstemeyer, Plasticity and fracture modeling/experimental study of a porous [79] R.W. Fonda, J.F. Bingert, Texture variations in an aluminum friction stir weld, Scr.
metal under various strain rates, temperatures, and stress states, J. Eng. Mater. Mater. 57 (11) (2007) 1052–1055, https://doi.org/10.1016/j.
Technol. 135 (4) (2013) 41008, https://doi.org/10.1115/1.4025292. scriptamat.2007.06.068.
[59] P.G. Allison, Y. Hammi, J.B. Jordon, M.F. Horstemeyer, Modeling and [80] F. Montheillet, M. Cohen, J.J. Jonas, Axial stresses and texture development during
experimental study of fatigue of powder metal steel (FC-0205), Powder Metall. 56 the torsion testing of Al, Cu and α-Fe, Acta Metall. 32 (11) (1984) 2077–2089,
(5) (2013) 388–396, https://doi.org/10.1179/1743290113Y.0000000063. https://doi.org/10.1016/0001-6160(84)90187-1.
[60] O.L. Rodriguez, P.G. Allison, W.R. Whittington, H. El Kadiri, O.G. Rivera, M. [81] A. Balasundaram, A.M. Gokhale, S. Graham, M.F. Horstemeyer, Three-dimensional
E. Barkey, Strain rate effect on the tension and compression stress-state asymmetry particle cracking damage development in an Al-Mg-base wrought alloy, Mater. Sci.
for electron beam additive manufactured Ti6Al4V, Mater. Sci. Eng. A 713 (2018) Eng. A 355 (1–2) (2003) 368–383, https://doi.org/10.1016/S0921-5093(03)
125–133, https://doi.org/10.1016/j.msea.2017.12.062. 00103-5.
[61] Y. Lang, Y. Cai, H. Cui, J. Zhang, Effect of strain-induced precipitation on the low [82] M. Lugo, J.B. Jordon, M.F. Horstemeyer, M.A. Tschopp, J. Harris, A.M. Gokhale,
angle grain boundary in AA7050 aluminum alloy, Mater. Des. 32 (8–9) (2011) Quantification of damage evolution in a 7075 aluminum alloy using an acoustic
4241–4246, https://doi.org/10.1016/j.matdes.2011.04.025. emission technique, Mater. Sci. Eng. A 528 (2011) 6708–6714, https://doi.org/
[62] C.B. Fuller, M.W. Mahoney, M. Calabrese, L. Micona, Evolution of microstructure 10.1016/j.msea.2011.05.017.
and mechanical properties in naturally aged 7050 and 7075 Al friction stir welds, [83] V. 2 ASM International Handbook, Heat treating, Metal Finishing, 101, p. 87, 2003,
doi: 10.1016/S0026-0576(03)90166-8.

17

You might also like