You are on page 1of 15

Mechanical behavior and strengthening mechanisms in ultrane

grain precipitation-strengthened aluminum alloy


Kaka Ma
a
, Haiming Wen
a,b
, Tao Hu
a
, Troy D. Topping
a
, Dieter Isheim
b,c
,
David N. Seidman
b,c
, Enrique J. Lavernia
a
, Julie M. Schoenung
a,
a
Department of Chemical Engineering and Materials Science, University of California Davis, One Shields Avenue, Davis, CA 95616, USA
b
Department of Materials Science and Engineering, Northwestern University, 2220 Campus Drive, Evanston, IL 60208-3109, USA
c
Northwestern University Center for Atom Probe Tomography (NUCAPT), 2220 Campus Drive, Evanston, IL 60208-3109, USA
Received 23 July 2013; received in revised form 18 September 2013; accepted 23 September 2013
Abstract
To provide insight into the relationships between precipitation phenomena, grain size and mechanical behavior in a complex precip-
itation-strengthened alloy system, Al 7075 alloy, a commonly used aluminum alloy, was selected as a model system in the present study.
Ultrane-grained (UFG) bulk materials were fabricated through cryomilling, degassing, hot isostatic pressing and extrusion, followed by
a subsequent heat treatment. The mechanical behavior and microstructure of the materials were analyzed and compared directly to the
coarse-grained (CG) counterpart. Three-dimensional atom-probe tomography was utilized to investigate the intermetallic precipitates
and oxide dispersoids formed in the as-extruded UFG material. UFG 7075 exhibits higher strength than the CG 7075 alloy for each
equivalent condition. After a T6 temper, the yield strength (YS) and ultimate tensile strength (UTS) of UFG 7075 achieved 734 and
774 MPa, respectively, which are 120 MPa higher than those of the CG equivalent. The strength of as-extruded UFG 7075 (YS:
583 MPa, UTS: 631 MPa) is even higher than that of commercial 7075-T6. More importantly, the strengthening mechanisms in each
material were established quantitatively for the rst time for this complex precipitation-strengthened system, accounting for grain-
boundary, dislocation, solid-solution, precipitation and oxide dispersoid strengthening contributions. Grain-boundary strengthening
was the predominant mechanism in as-extruded UFG 7075, contributing a strength increment estimated to be 242 MPa, whereas Oro-
wan precipitation strengthening was predominant in the as-extruded CG 7075 (102 MPa) and in the T6-tempered materials, and was
estimated to contribute 472 and 414 MPa for CG-T6 and UFG-T6, respectively.
2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Al alloys; Precipitation; Strengthening mechanism; Ultrane-grained materials; Atom-probe tomography
1. Introduction
Conventional coarse-grained (CG) precipitation-
strengthened 7000 series aluminum (Al) alloys have been
widely used for aerospace and transportation applications
because of their high strength and heat treatability [1,2].
In this class of precipitation-strengthened alloys, extremely
small and uniformly dispersed precipitates, which act as
obstacles to dislocation movement, form within the Al
matrix upon heat treatment and thus strengthen the mate-
rials [3]. This phenomenon is generally referred to as pre-
cipitation strengthening. Natural aging, or precipitate
formation at room temperature, occurs in most 7000 series
alloys [4,5]. It has been generally accepted that the precip-
itation starts with the formation of GuinierPreston (GP)
zones, which may be regarded as coherent metastable pre-
cipitates [3,68]. The typical diameter of GP zones is of the
order of a few nanometers [8]. Subsequent evolution of the
precipitates involves the replacement of the metastable GP
zones with a metastable semicoherent phase, g
0
-MgZn
2
[9].
This occurs primarily because GP zones are isostructural
1359-6454/$36.00 2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.actamat.2013.09.042

Corresponding author. Tel.: +1 (530) 752 5840; fax: +1 (530) 752 9554.
E-mail address: jmschoenung@ucdavis.edu (J.M. Schoenung).
www.elsevier.com/locate/actamat
Available online at www.sciencedirect.com
ScienceDirect
Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
with the matrix and, therefore, have a lower interfacial
energy than intermediate or equilibrium precipitate phases
that possess a dierent crystal structure. As a result, the
nucleation barrier for GP zones is signicantly smaller
[1]. The incoherent equilibrium hexagonal phase, g-
MgZn
2
, forms from g
0
-phase precipitates at higher aging
temperatures and longer aging times. These g-phase pre-
cipitates are generally larger in size (diameter > 50 nm)
and are preferentially located at grain boundaries. The pre-
cipitation sequence can be summarized as follows: super-
saturated solid solution !GP zones !g
0
(MgZn
2
) !g
(MgZn
2
) [3,8]. Basically, the presence of a high number
density of GP zones and ne g
0
-phase precipitates is
responsible for the strengthening of the material [6,10]. In
an eort to accelerate aging kinetics, articial aging is per-
formed at higher temperatures, during which the strength
achieves a maximum value. After long aging times or
higher aging temperatures, the strength begins to decrease
and the alloy becomes over-aged [3,11]. The commonly
used articial aging treatment, called a T6 temper, that
results in peak microhardness for these alloys with conven-
tional coarse grains, is performed at 120 C for 24 h [5].
Interest in nanocrystalline or nanostructured (NS; grain
diameter < 100 nm) and ultrane-grained (UFG; grain
diameter > 100 nm, but less than 1000 nm) materials, orig-
inally motivated by reports of novel deformation mecha-
nisms as well as by the potential to attain notable
enhancements in mechanical properties [1216], has gradu-
ally moved from pure metals and simple alloys to more
complex precipitation-strengthened alloys with many alloy-
ing components. Prior studies of Al 7075, a representative
precipitation-hardenable aluminum alloy, have revealed
that these alloys can be further strengthened by incorporat-
ing grain renement. Zhao et al. [4,9] fabricated UFG 7075
with a grain diameter of 400 nm utilizing commercial
7075 rod (grain diameter 40 lm) through equal channel
angular pressing (ECAP). They reported that the yield
strength and the tensile strength of the UFG 7075 were
650 and 720 MPa, respectively, with natural aging for a
month after ECAP, which represent strength increases of
103% and 35%, respectively, over its commercial Al 7075
counterpart. The improvement in the strength was ascribed
to grain renement and higher number densities of both
GP zones and dislocations in the UFG material. In a
related study, Zhao et al. also documented a simultaneous
increase in both ductility and strength for NS 7075 (aver-
age grain diameter 100 nm, processed by cryorolling)
with subsequent articial aging compared with the unaged
condition [17]. In addition to the ne GP zones, g
0
- and g-
phase precipitates were introduced in the nanograins
through aging, increasing the dislocation density. The
increased dislocation density led to an improvement in
the work-hardening rate and consequently contributed to
the enhanced uniform elongation. It was concluded that
the high dislocation density and ne grain size of the NS
sample were primarily responsible for its improved strength
over the CG sample, while the high density of second-phase
precipitates was responsible for its improved ductility over
as-processed NS 7075 without aging [17]. In a related
study, Panigrahi and Jayaganthan [18] applied cryorolling
to produce an UFG 7075 material with high-angle grain
boundaries that exhibited improved strength due to the
HallPetch eect and a higher dislocation density. It was
documented that the microhardness and tensile strength
of the cryorolled UFG 7075 was reduced after annealing
at temperatures of 150250 C and subsequently remained
constant when the annealing temperature was increased
[19]. More recently, a NS 7075 material exhibiting an extre-
mely high yield strength of 1 GPa, combined with a uni-
form elongation of 5%, was successfully produced by
high-pressure torsion (HPT) [20]. It was suggested that
the formation of a nanostructured architecture, which
comprised a solid solution including a high number density
of dislocations, sub-nanometer intragranular solute clus-
ters, nanometer-scale intergranular solute clusters and
grains of tens of nanometers in diameter, contributed to
the dramatic increase in strength.
Despite ample evidence that grain renement further
improves the strength of precipitation-strengthened Al
alloys, precise determination of the underlying strengthen-
ing mechanisms has been hindered by the complexity of the
possible mechanisms, including: grain-boundary strength-
ening (HallPetch eect), solid-solution strengthening, dis-
location strengthening and precipitation strengthening.
Accordingly, the goal of the present study is to formulate
a quantitative insight into strengthening mechanisms by
providing a direct comparison of an UFG precipitation-
strengthened material with an otherwise equivalent powder
metallurgy (PM)-derived CG material. It is noted that
unlike CG materials made by casting, the PM-derived
CG material is expected to exhibit a relatively ne grain
diameter, in the range of 15 lm [21]. To the best of our
knowledge, this is the rst time a direct comparison
between UFG and CG materials, consolidated and heat-
treated using identical processing steps, has been docu-
mented for such a complex precipitation-strengthened Al
alloy. More importantly, fundamental insights into the
interrelationships between grain renement, precipitation
characteristics and mechanical behavior are provided.
2. Experimental procedure
Al 7075 was chosen as a representative alloy for this
investigation, partly because the precipitation sequence
and kinetics in this system have been extensively studied.
Cryomilling, a mechanical attrition technique in a cryo-
genic environment [14,15,2224], was utilized in our study
to obtain NS 7075 powder. This technique takes advantage
of the low boiling temperature, 77.2 K, of liquid nitrogen,
which suppresses recovery and recrystallization in the pow-
der, and leads to nanocrystalline grain structures and rapid
grain renement. Cryomilling also results in a high number
density of dislocations in the material through severe plas-
tic deformation (SPD) [14,15]. Additionally, ne oxide/
2 K. Ma et al. / Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
nitride dispersoids are introduced into the material during
cryomilling [25,26]. UFG 7075 bulk materials were subse-
quently fabricated by consolidating the nanocrystalline
powder through degassing, hot isostatic pressing, and
extrusion [27]. Equivalent CG 7075 bulk materials were
fabricated employing the same techniques by consolidating
gas atomized powder. The microstructures of the samples
were characterized by transmission electron microscopy
(TEM) and X-ray diraction (XRD). Atom-probe tomog-
raphy (APT) [2832] was utilized to characterize the inter-
metallic precipitates and oxide dispersoids in the UFG
7075 on an atomic scale; this advanced technique deter-
mines the elemental identity and position of individual
atoms with sub-nanometer resolution in three dimensions
(3-D), and therefore enables the quantication of the size,
morphology, composition, number density and volume
fraction of the nanoscale precipitates and dispersoids [28
30,3335]. The mechanical behavior of the materials was
investigated through standard tensile testing. Details are
provided below:
2.1. Material processing
2.1.1. Powder modication
The starting material used in the present study is gas-
atomized Al 7075 powder provided by Valimet, Inc.
(Stockton, CA) with a particle diameter of 200 mesh
(74 lm). The nominal chemical composition of the pow-
der is provided in Table 1 [36]. The feedstock powder was
cryomilled in liquid nitrogen (LN
2
) for 12 h in a modied
Szegvari attritor. The cryomilling parameters are summa-
rized in Table 2. The cryomilled powder was subsequently
lled into a 1 inch diameter by 3 inch long can fabricated
from Al 6061 [36] and hot vacuum degassed at tempera-
tures up to 500 C for 12 h with a nal pressure in the range
10
6
Torr [37]. Degassing is necessary to remove hydrates
and stearates attributable to powder handling, and the pro-
cess control agent (PCA), stearic acid, utilized during cryo-
milling [37].
2.1.2. Consolidation
After cryomilling and hot vacuum degassing, the can
was crimped, welded and hot isostatically pressed (HIPed)
at 400 C and 172 MPa (25 ksi) in argon to achieve full
consolidation. The Al 6061 can material was removed by
machining and the consolidated Al 7075 was machined into
a workable billet, which subsequently underwent a slow-
strain-rate (SSR) extrusion at 350 C using a 3.56 MN
(400 ton) Hypress Technologies Inc. model UGP400-6
press equipped with a custom-built resistive band heating
system. Omega temperature controllers using K-type ther-
mocouples measured and controlled the temperature of the
die. The extrusion die is a cylinder of 108 mm
(4.25 inches) diameter fabricated from a H13 tool steel with
a 25.4 mm (1 inch) hole bored through its center. The
reduction ratio was 10:1 in area ratio, which gave a nal
diameter of 7.9 mm. The strain rate, 3.7 10
2
s
1
,
was based on a nominal velocity of 100 mm min
1
. For
the CG 7075 counterpart, as-received gas-atomized Al
7075 powder was degassed and consolidated following
the same processing steps as used for the cryomilled
powder.
2.1.3. Heat treatment
Solution heat treatment was performed on some of the
tensile specimens, which were wrapped in Al foil, in an elec-
tric resistance furnace (model C601K, Cress Mfg Co., El
Monte, CA) at 500 C for 1 h. Subsequently, the specimens
were quenched in ice water. The articial aging treatment
was performed at 120 C in a tube furnace (Carbolite,
UK) for 24 h. The schedule for the articial heat treatment
was chosen following the T6 temper for conventional Al
7075 materials [5]. Natural aging occurred after solution-
ized samples were stored at room temperature for a week.
A summary of the sample identication (ID) and the corre-
sponding process/heat-treatment conditions is provided in
Table 3. A pictorial description of the thermal history the
materials experienced is provided in Fig. 1 [38].
2.2. Characterization
2.2.1. Room temperature tensile testing
Cylindrical, threaded tensile specimens were machined
with a gauge length of 12 mm (0.5 inch) 3 mm
(0.12 inch) diameter, with dimensions close to the specica-
tions for sub-size ASTM E 8M standard. The tensile tests
were performed at room temperature using a universal test-
ing machine (Instron 8801, Norwood, MA) with strain
measured by a standard video extensometer. The strain
rate utilized was 10
3
s
1
. Two samples were tested for
each condition to conrm reproducibility. The solution
Table 1
Nominal chemical composition of Al 7075 alloys [36].
Elements Al Zn Mg Cu Cr Fe Mn Si Ti Other
wt.% Balance 5.16.1 2.12.9 1.22 0.180.28 60.5 60.3 60.4 60.2 60.15
Table 2
Parameters used in cryomilling.
Cryomilling parameters
Milling media Slurry of stainless steel balls and LN
2
Ball-to-powder weight
ratio
32:1
Impellor rotation speed 180 rpm
Milling time 12 h
Milling temperature 183 to 190 C
Process control agent
(PCA)
Stearic acid 2 g (0.2 wt.% of the loaded
powder)
K. Ma et al. / Acta Materialia xxx (2013) xxxxxx 3
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
heat-treated samples and T6-tempered samples were all
tested within 1 h after the heat treatment to exclude any
natural aging eect.
2.2.2. Phase identication and microstructure
XRD analyses were performed on both the powder and
consolidated bulk samples to investigate the phase consti-
tution after each processing step, using a Scintag X-ray dif-
fractometer equipped with a graphite monochromator
using Cu K
a
(k = 0.15406 nm) radiation. Additionally,
quantitative analyses of grain diameter and microstrain
were performed on the NC7075 powder and UFG7075-E
consolidated sample according to the WilliamsonHall
method [39]. As a measure of peak broadening, the full-
width at half-maximum (FWHM) of the peaks was
obtained by tting the XRD peak proles using the Pear-
son 7 function. The true peak broadening, B, was derived
from:
B

B
2
obs
B
2
inst
q
; 1
where B
obs
is the observed peak broadening and B
inst
is the
instrumental broadening.
The microstructures of the powder and consolidated
samples were characterized employing a Phillips CM12
transmission electron microscope and a JEOL 2500 high-
resolution transmission electron microscope, operating at
120 and 200 kV, respectively. The grain diameter was esti-
mated by measuring and averaging the length and width of
the strongly diracting grains using ImageJ

image analy-
sis software [40]. Truncation eects caused by the surface of
the TEM foil were neglected. APT studies of the as-
extruded UFG 7075 were performed using a Cameca
local-electrode atom-probe (LEAP) 4000X-Si tomograph
[2831]. Parallelepipeds of 400 lm 400 lm 8 mm were
cut and subsequently electropolished at ambient tempera-
ture to obtain a needle-shaped tip with a radius of curva-
ture of 50 nm at the apex. APT was performed using
UV picosecond laser pulsing (355 nm wavelength), a pulse
repetition rate of 500 kHz, a pulse energy of 200 pJ and a
detection rate of 0.20.5%. The specimen base temperature
was 60 K, and steady-state DC voltages between 2.0 and
6.0 kV were applied for controlled eld-evaporation in
the voltage pulsing mode. 3-D reconstructions and data
analyses of the APT data were performed using Camecas
IVAS

software, version 3.6.1. Compositions of precipi-


tates and dispersoids were obtained employing the proxim-
ity histogram concentration prole technique [41]. More
details regarding APT can be found in Refs. [25,2831].
3. Results
3.1. Tensile behavior
The representative engineering tensile stressstrain
curves for CG and UFG materials for dierent heat treat-
ment conditions are presented in Fig. 2a and b, respec-
tively. As a reference, a commercial extruded 7075 bar,
T6 temper, was tested in the same condition as the other
samples; the result is presented as curve 5. The mechanical
property data are summarized in Table 4. It is clear that the
UFG 7075 materials exhibit higher strengths than the CG
7075 materials for each equivalent condition. In the
extruded condition, the tensile yield strength, r
ys
, and ulti-
mate strength, r
uts
, of sample UFG7075-E are 583 and
631 MPa, respectively, which are approximately 106%
and 45% greater, respectively, than those of sample
CG7075-E (283 and 436 MPa). For the T6 temper, the
Table 3
Sample identication (ID) and the corresponding processing conditions.
Sample ID Processing conditions
NC7075 As-cryomilled 7075 powder
UFG7075-E Consolidated 7075 made from NC7075, which went
through degassing plus HIP plus extrusion
UFG7075-E-sol UFG7075-E was solution heat treated and quenched
in ice water
UFG7075-E-Nag UFG7075-E-sol was naturally aged at ambient
temperature in air for 1 week
UFG7075-E-T6 UFG7075-E-sol was articially aged at 120 C in air
for 24 h
CG7075-E Consolidated 7075 made from as-received gas
atomized (i.e. coarse grained, CG) powder, which
went through degassing plus HIP plus extrusion
CG7075-E-sol CG7075-E was solution heat treated and quenched in
ice water
CG7075-E-Nag CG7075-E-sol was naturally aged at ambient
temperature in air for 1 week
CG7075-E-T6 CG7075-E-sol was articially aged at 120 C in air for
24 h
C7075-T6 Commercial extruded 7075 bar in T6 temper
Fig. 1. Temperature vs. processing time for Al 7075 alloy from powder to consolidated UFG bulk samples and subsequent T6 temper [38].
4 K. Ma et al. / Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
r
ys
and r
uts
values of sample UFG7075-E-T6 increased to
734 and 774 MPa, respectively, which are 20% and 17%,
respectively, greater than those of sample CG7075-E-T6
(613 and 659 MPa). Moreover, the r
ys
and r
uts
values of
sample UFG7075-E-T6 are 36% and 31% greater, respec-
tively, than those of the commercial Al 7075-T6 alloy
(541 and 593 MPa). Note that the strength of the as-
extruded UFG7075-E sample is also greater than that of
the commercial Al 7075-T6 alloy and is close in value to
that of sample CG7075-E-T6. These results indicate that
the cryomilling process might eliminate the need for the
solutionizing and aging treatments to achieve the desirable
strength, with acceptable ductility (elongation of 2.8%).
Additionally, a decrease in stress, i.e. strain softening,
occurred after yielding in all the curves for the UFG
7075 materials, while this phenomenon was absent for the
CG 7075 materials. Strain softening has been observed in
other Al alloys, e.g. Al 5083, that were cryomilled and con-
solidated in a similar manner [24,27,42]. This phenomenon
is, however, beyond the scope of the current study.
Interestingly, after the solutionizing treatment, the CG
7075 material exhibited a decrease in r
ys
from 283 MPa
(CG7075-E, curve 1) to 236 MPa (CG7075-E-sol, curve
2) with almost the same ductility; while sample
UFG7075-E-sol (curve 7) maintained both the same r
ys
and r
uts
as those of sample UFG7075-E (curve 6), but
exhibited an increase in elongation from 1.9% to 5.3%.
Additionally, both natural aging and articial aging con-
tributed more signicantly to increasing the strength in
the CG 7075 materials than in the UFG 7075 materials.
The increases in r
ys
for samples CG7075-E-Nag and
CG7075-E-T6 are 70% and 160%, respectively, compared
with the baseline values for sample CG7075-E-sol. In con-
trast, the increases in r
ys
after natural aging and T6 temper
of the UFG 7075 materials are only 13% and 26%,
respectively.
3.2. Microstructure
3.2.1. Grain diameter
Fig. 3 displays a representative bright-eld TEM image
of the as-cryomilled Al 7075 powder (NC7075) and a histo-
gram of the grain diameter distribution that ranges from 20
to 90 nm with a mean value of 46 nm, which demonstrates
that nanocrystalline Al 7075 powder was successfully fabri-
cated during the cryomilling process. No contrast related
to precipitates could be discerned in the TEM micrographs
and we concluded that the alloy is largely precipitation-free
after the cryomilling process. Due to the thermal exposure
during degassing, HIPing and extrusion, signicant grain
growth occurred in the materials, as observed in the micro-
graphs in Fig. 4. The length of the elongated grains in the
UFG7075-E sample ranged from 95 to 730 nm along the
extrusion direction (Fig. 4a) [38]. Averaging the mean
length (315 nm) and the mean width (175 nm) of the elon-
gated grains yields an estimate for the mean grain diameter
of 245 nm [38]. In contrast, sample CG7075-E exhibits
much coarser equiaxed grains ranging from 0.5 to 1 lm,
with an average grain diameter of 894 nm (Fig. 4b). After
the T6 temper, the average grain diameter of the UFG
7075 increased to 422 nm (Fig. 4c), while that of the CG
7075 increased to about 1 lm (Fig. 4d). Additional TEM
images and grain diameter histograms appear elsewhere
[38].
3.2.2. Precipitates
A detailed and robust investigation of the precipitates in
terms of type, morphology, size and density, as well as for-
mation mechanisms, in both the CG and UFG 7075 mate-
rials, in both the as-fabricated condition and after a T6
temper, can be found elsewhere [38]. The morphology, size
and number density of the three families of precipitates
GP zones, and g
0
- and g-phase precipitateswere
observed to vary signicantly between the CG and UFG
samples. The critical relevant results for explaining the
Fig. 2. Comparison of tensile stressstrain curves for: (a) CG 7075
materials and (b) UFG 7075 materials, after selected heat-treatment
conditions. Curve 5 is for a reference specimen of commercially extruded
7075 bar in T6 temper, which was tested under the same conditions as the
other samples. Symbols are added every 50 data points to dierentiate the
curves from one another.
K. Ma et al. / Acta Materialia xxx (2013) xxxxxx 5
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
mechanical behavior are summarized in Table 5. Represen-
tative TEM bright-eld images of the consolidated CG and
UFG 7075 materials are displayed in Fig. 4. Additional
TEM images, found in a prior study [38], support the
description of the precipitates provided herein. Sample
UFG7075-E possessed a relatively low number density of
GP zones and g
0
-phase precipitates, both with average
diameters <5 nm, as observed within the grain interiors
[38]. Irregularly shaped g-phase MgZn
2
precipitates, which
contain some Cu and Al in solution with the Zn, were occa-
sionally observed, with an average diameter of 90 nm [38].
In contrast, sample CG7075-E (Fig. 4b) exhibited a num-
ber of plate-like g
0
-phase (diameter 55 nm) and lath-like
g-phase precipitates (length 174 nm) in grain interiors;
these precipitates were located along dislocation lines
[38]. After the T6 temper, the grains in the UFG 7075
material (Fig. 4e) contained a large number of nanoscale
spherical GP zones (average diameter 3 nm) and plate-
let-like g
0
-phase precipitates (diameter 6 5 nm); in con-
trast, although sample CG7075-E-T6 (Fig. 4f) also
exhibited many spherical GP zones, with diameter <5 nm,
the g
0
-phase precipitates were coarser with an average
diameter of 61 nm [38].
To incorporate this descriptive information concerning
the precipitates into the strengthening mechanism estimates
(Section 4.4) it was also necessary to estimate the mean
radii and mean edge-to-edge interprecipitate spacing, k
p
.
The mean radii, r, in Table 5, were estimated by averaging
the radii of all the precipitates regardless of precipitate
morphology. The values for k
p
, in Table 5, were estimated
from 60 measurements on several representative TEM
images.
3.2.3. X-ray diraction
The XRD patterns for the UFG and CG 7075 materials
are displayed in Fig. 5a and b, respectively. The g
0
- and g-
phases were evaluated together due to the overlapping of
their XRD peaks. The XRD peaks for the g
0
- and g-phases
are present in the as-received gas-atomized 7075 powder,
but disappear after cryomilling, indicating that cryomilling
dissolved the precipitates and solutionized the powder to
within the limits of detection. The g
0
- and g-phases were
detected, however, in the consolidated samples and the
intensity of these XRD peaks are stronger for sample
CG7075-E than for sample UFG7075-E, which is consis-
tent with our TEM observations. During the solutionizing
step, the g-phase precipitates were dissolved in samples
CG7075-E-T6 and UFG7075-E-T6. The amount of the
g
0
-phase was reduced, while the number density of GP
zones increased after T6 temper in sample CG7075-E-T6.
Therefore, the XRD peaks for the g
0
- and g-phases became
weaker for sample CG7075-E-T6 compared to those for
sample CG7075-E. Similarly, the dissolution of the
g-phase precipitates and the extremely small size of the
g
0
-phase precipitates resulted in a diminution of their
XRD peaks in sample UFG7075-E-T6 compared to sample
UFG7075-E.
The grain diameter, d, and microstrain, e, of the bulk
samples were calculated from the XRD peak broadening,
B, using the WilliamsonHall method [39]. This approach
assumes that B consists of grain renement broadening
and strain broadening, which is given by:
Bcos h
B

Kk
d
e sin h
B
; 2
Table 4
Mechanical data for the consolidated Al 7075 materials.
Sample ID 0.2%YS r
ys
(MPa) UTS r
uts
(MPa) Strain at fracture (%) Elongation (%)
CG7075-E 283 436 11.9 11.4
UFG7075-E 583 631 3.0 1.9
CG7075-E-sol 236 441 11.9 11.3
UFG7075-E-sol 584 635 6.4 5.3
CG7075-E-Nag 402 564 13.6 11.5
UFG7075-E-Nag 658 686 3.5 2.5
CG7075-E-T6 613 659 5.0 3.9
UFG7075-E-T6 734 774 4.0 2.8
C7075-T6 541 593 14.1 13.2
Fig. 3. Representative TEM image of as-cryomilled Al 7075 powder
(NC7075) with an embedded grain size histogram.
6 K. Ma et al. / Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
Fig. 4. Representative TEM images of samples: (a) UFG7075-E; (b) CG7075-E; (c) UFG7075-E-T6; (d) CG7075-E-T6; (e) a higher-magnication image
of UFG7075-E-T6; and (f) a higher-magnication image of CG7075-E-T6.
Table 5
Summary of the type, average diameter, mean radius and mean edge-to-edge interprecipitate spacing of the precipitates present in select samples.
Sample ID Precipitate type
[38]
Average diameter 2r (nm)
[38]
Mean radius of all the precipitates r
(nm)
Mean edge-to-edge interprecipitate spacing
k
p
(nm)
UFG7075-E GP zone 2.5 1.5 173
g
0
3.3
UFG7075-E -
T6
GP zone 3.1 1.5 19
g
0
4.4 (length)
1.5 (width)
CG7075-E g
0
55 40 194
g 174 (length)
35 (width)
CG7075-E-T6 GP zone 2.9 16 35
g
0
61
K. Ma et al. / Acta Materialia xxx (2013) xxxxxx 7
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
where k is the wavelength of Cu K
a
radiation, i.e. 1.54 A

, K
is 0.9, e is the microstrain and h
B
is the Bragg angle [25].
Plotting B cos h
B
vs. sin h
B
and performing a linear regres-
sion analysis, the values of d and e were obtained from the
slope and intercept of the tted curve. For materials sub-
jected to SPD, the dislocation density, q, in terms of d
and e is given by [4]:
q
2

3
p
e
db
; 3
where b = 0.286 nm is the magnitude of the Burgers vector
for Al [43]. Applying the values of d and e obtained from
Eq. (2), the values of q for samples UFG7075-E,
UFG7075-E-T6, CG7075-E and CG7075-E-T6 are
4.5 10
14
, 4.1 10
14
, 1.7 10
14
and 5.6 10
13
m
2
,
respectively. These data are used to calculate the contribu-
tion to strengthening from dislocations.
3.2.4. Atom-probe tomography
An APT tomographic 3-D reconstruction of a volume in
the UFG7075-E sample is presented in Fig. 6ad. Images
of each individual element (Zn, Mg, O or N) are displayed
in a separate box, with 8 at.% Zn, 8 at.% Mg, 5 at.% O and
5 at.% N isoconcentration surfaces superimposed, respec-
tively; this volume contains 23 million atoms. The precipi-
tates are clearly delineated by the isoconcentration
surfaces. The inhomogeneous distributions of the Zn,
Mg, O and N atoms indicate the formation of precipitates
in the material. The locations of Mg-enriched regions coin-
cide with Zn- or O-enriched region, indicating the coprecip-
itation of Mg and Zn atoms, and the presence of MgO
dipsersoids. A slight enrichment of N is observed where
O is enriched, suggesting that some MgO dispersoids con-
tain a small concentration of N. The presence of N is a
direct result of cryomilling, which occurs in liquid nitrogen
[14,22,44]. Proximity histogram concentration proles [41]
were employed to quantify the specic chemical composi-
tions of the precipitates and dispersoids. Results from a
number of proximity histograms reveal that three distinct
types of precipitates and dispersoids exist in sample
UFG7075-E: MgZn GP zones, MgZn
2
g
0
-phase precipi-
tates and MgO dispersoids. Fig. 7a displays isoconcentra-
tion surfaces of 8 at.% Zn with a representative example
of a GP zone indicated by an arrow; the corresponding rep-
Fig. 5. X-ray diraction patterns for (a) UFG 7075 and (b) CG 7075
materials.
Fig. 6. Atom-probe tomographic 3-D reconstructions of sample
UFG7075-E: (a) reconstruction with only Zn atoms displayed and a
superimposed 8 at.% Zn isoconcentration surface; (b) reconstruction with
only Mg atoms displayed and a superimposed 8 at.% Mg isoconcentration
surface superimposed; (c) reconstruction with only O atoms displayed and
a superimposed 5 at.% O isoconcentration surface; and (d) reconstruction
with only N atoms displayed and a superimposed 5 at.% N
isoconcentration.
8 K. Ma et al. / Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
resentative proximity histogram is in Fig. 7b. Examples of
g
0
-phase precipitates and MgO dispersoids and the corre-
sponding proximity histograms are displayed in Fig. 7cf,
respectively. Although some of the GP zones and g
0
-phase
precipitates contain up to 5 at.% Cu, and some of the
MgO dispersoids contain up to 10 at.% N, these phases
are referred to below without reference to their complex
chemistries. Quantitative information on precipitate diam-
eter distribution, number density and volume fraction were
also obtained following the methodology detailed in Refs.
[25,35]. The precipitate diameter distributions for the GP
zones, g
0
-phase precipitates and MgO dispersoids are pre-
sented in Fig. 8a and b, respectively. The standard devia-
tion was used as the measure of uncertainty in the
precipitate diameters. The statistical precipitate and disper-
soid characteristics are presented in Table 6. The average
diameter of GP zones/g
0
-phase precipitates is 3.5 nm,
which agrees with the TEM observations. The volume frac-
tion of these precipitates is about 0.07%, and their number
density is 1.8 10
22
m
3
. Although TEM was unable to
reveal the presence of MgO dispersoids, the APT results
indicate an average diameter of 4 nm, with a volume frac-
tion of 0.14% at a number density of 1.3 10
22
m
3
.
4. Discussion
Our recent investigation [38] revealed that grain size
inuences precipitation kinetics, resulting in signicant dif-
ferences in the size, composition and spatial distribution of
precipitates between the CG and UFG 7075 materials. GP
zones, as well as plate-like g
0
-phase precipitates, were
observed to nucleate homogeneously in the grain interior
of sample UFG7075-E; in contrast, large numbers of
g
0
-phase precipitates formed either on dislocation lines or
in the vicinity of the dislocations in sample CG7075-E.
During articial aging, a high number density of GP zones
with an average diameter of 3 nm and platelet-shaped
g
0
-phase precipitates formed via homogeneous nucleation
and growth in the interior of the grains in the UFG mate-
rial. Alternatively, the presence of dislocations in the grains
of the CG material assisted the heterogeneous nucleation
and growth of plate-like g
0
-phase precipitates, whereas
Fig. 7. (a) 8 at.% Zn isoconcentration surface for sample UFG7075-E, where the isoconcentration surface of a precipitate used for a proximity histogram
analysis is highlighted and arrowed; (b) proximity histogram concentration prole based on the 8 at.% Zn isoconcentration surface of the precipitate
arrowed in (a); (c) 8 at.% Zn isoconcentration surface, where the isoconcentration surface of a precipitate used for the proximity histogram analysis is
highlighted and arrowed; (d) proximity histogram concentration prole based on the 8 at.% Zn isoconcentration surface of the precipitate arrowed in (c);
(e) 5 at.% O isoconcentration surface, where the isoconcentration surface of a precipitate used for proximity histogram analysis is highlighted and arrowed;
and (f) proximity histogram concentration prole based on the 5 at.% O isoconcentration surface of the precipitate arrowed in (e).
K. Ma et al. / Acta Materialia xxx (2013) xxxxxx 9
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
GP zones formed homogeneously at a high number density
in regions without dislocations in sample CG7075-E-T6
[38]. It was proposed that a reduction in grain size, with
an increase in grain boundary area per unit volume,
resulted in a small concentration of vacancies in the
UFG 7075, and thus homogeneous nucleation of the pre-
cipitates was inhibited during aging [38]. The dislocation
substructure, which provided heterogeneous nucleation
sites for g
0
-phase precipitates, governed the precipitation
kinetics in CG 7075 [38]. The modication of the feedstock
powder by cryomilling reduced the length scale of the grain
size and subsequently changed the morphology, size and
nucleation mechanism of the precipitates in the nal con-
solidated UFG 7075 material, as well as introducing nano-
scale oxide dispersoids. The current investigation
demonstrates that these microstructural changes conse-
quently lead to an increase in tensile strength in the UFG
7075 when compared with CG 7075. To provide insight
into the measured dierences in the stressstrain behavior
between these materials, it is important to establish the
active strengthening mechanisms. The CG and UFG
7075 materials have experienced several thermomechanical
processing (TMP) steps during the consolidation proce-
dure, in which signicant plastic deformation was also
introduced. Considering the nature of this class of precipi-
tation-strengthened Al alloys, the inuence of cryomilling
and the plastic deformation that occurred during TMP,
the following strengthening mechanisms need to be evalu-
ated to explain the higher strength in the UFG 7075 mate-
rials: grain-boundary strengthening, solid-solution
strengthening, dislocation strengthening and precipitate/
dispersoid strengthening. Quantitative estimates of the con-
tribution of each mechanism in the UFG and CG 7075
materials are described below; the strength increments
stemming from the four mechanisms are summarized in
Table 7. The base strength of pure Al, the lattice friction
stress, r
0
, is not included in this table. Although there are
various studies in the literature in which linear summation
of strength increments have been applied [25,35,45], this
was not the purpose of the current investigation. Rather,
our goal was to highlight the dierences in mechanisms
that dominate strengthening behavior between the UFG
and CG materials. The physical meaning and values of
the symbols in the equations used in the current study
are summarized in Table 8 [25,36,43,4650].
4.1. Grain-boundary strengthening (HallPetch eect)
One of the most signicant consequences of cryomilling
is grain size renement, creating a high volumetric density
of grain boundaries that impede dislocation movement and
dislocation propagation to adjacent grains, thereby
strengthening the materials [14,15,51]. The grain-boundary
strengthening mechanism is usually described by the Hall
Petch equation [47,5254]:
r
y
r
0

k
y

d
p ; 4
where d is the average grain diameter, r
0
is the friction
stress and k
y
is the HallPetch slope.
Several publications have reported a breakdown or devi-
ation in the HallPetch relationship in UFG materials fab-
ricated by SPD processes such as accumulative roll
bonding (ARB) and HPT [5557]. Two explanations have
been proposed for this breakdown: (i) the eective grain
size for the HallPetch relationship is larger than the mea-
sured value because the mobile lattice dislocations pass eas-
ily through the non-equilibrium grain boundaries that were
generated in the material during the SPD processing [55];
Fig. 8. Precipitate diameter distributions for: (a) GP zones/g
0
-phase and
(b) MgO. The uncertainty in the average precipitate diameter corresponds
to the standard deviation.
Table 6
Summary of precipitate characteristics in sample UFG-7075-E from atom-
probe tomography (APT).
Precipitates Average diameter
(nm)
Volume
fraction (%)
Number density
(m
3
)
GP zone/g
0
-
MgZn
2
3.5 1.6 0.07 1.8 0.5 10
22
MgO 4.0 3.1 0.14 1.3 0.4 10
22
10 K. Ma et al. / Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
(ii) increased participation of extrinsic dislocations are able
to move in the non-equilibrium grain boundaries and
thereby reduce the apparent strength [57]. Non-equilibrium
grain boundaries are responsible for the deviation from a
HallPetch behavior. In our work, the non-equilibrium
grain boundaries introduced during cryomilling are eec-
tively eliminated due to the annealing that occurs during
the subsequent degassing step and thermomechanical con-
solidation steps, which involve prolonged thermal exposure
at elevated temperatures (see Fig. 1). This is supported by
published studies [14,15,58], which fail to reveal non-equi-
librium grain boundaries in the microstructure of consoli-
dated cryomilled materials. Also, several published
studies have applied Dr = kd
1/2
to estimate grain-bound-
ary strengthening for consolidated cryomilled materials
with grain size ranges similar to those in the current work
[45,58,59]. Therefore, it is reasonable to assume that a
HallPetch relationship with a power of d
1/2
is applicable
to the UFG Al 7075 in the current study. Wert et al. stud-
ied the eect of grain size on the yield strength of 7000 ser-
ies Al alloys and revealed that the HallPetch coecient,
k
y
, for peak-aged Al 7075 was 0.12 MPa=

m
p
[48,49].
Assuming the values of r
0
are the same for CG and
UFG 7075, the strength increase from grain-boundary
strengthening, Dr
gb
, is proportional to d
1/2
. For UFG
7075, the grain diameter was obtained by averaging the
length and the width of the elongated grains. Therefore,
the increase in yield strength due to grain-boundary
strengthening is calculated to be 242, 185, 127 and
120 MPa for samples UFG7075-E, UFG7075-E-T6,
CG7075-E and CG7075-E-T6, respectively. It is evident
that the strength increment from grain-boundary strength-
ening in sample UFG7075-E is almost twice that in sample
CG7075-E, while the strength increment from grain-
boundary strengthening in sample UFG7075-E-T6 is 54%
higher than in sample CG7075-E-T6.
4.2. Solid-solution strengthening
Solid-solution strengthening occurs when other elements
are alloyed with a metal matrix as solute atoms that dier
from the matrix atoms in size and/or shear modulus, which
can cause a variation of strain elds. Local strain elds are
created that interact with dislocations and impede their
motion, leading to an increase in the yield strength of the
material. It has been generally accepted that solid-solution
strengthening is governed by the Fleischer equation [60,61]:
Dr
ss
MGbe
3
2
ss

c
p
: 5
The meaning and values of the symbols in this equation
are listed in Table 8. Some studies have demonstrated a need
to modify the power of c from to 1 for nanostructured
materials with grain diameters <30 nm [62]. Nevertheless,
this is not necessary for the present systembecause the grain
size in the UFG7075 is of the order 100 nm and larger. The
Fleischer equation has been widely used in the literature
for UFG materials with grain diameters in this range
[25,35,45]. Therefore, it is reasonable to assume that Eq.
(5) is applicable in the current investigation. From Eq. (5),
the value of Dr
ss
depends on the dierence in the shear
moduli between the solute and the matrix, the concentration,
c, and the dierence in size between the solute and solvent
atoms (causing lattice strain, e). Table 1 demonstrates that
Al 7075 primarily contains Zn, Mg and Cu solute atoms.
The dierence in radii and the theoretical contributions to
the yield strength from these elements are listed in Table 9,
with data for high-purity binary solid-solution alloys
[50,63]. Mg, Zn and Cu are not all in solid solution in the
extruded condition nor in the T6-tempered condition
because they form second-phase precipitates or segregate
to the grain boundaries [38]. As an upper bound, assuming
that all of the solute atoms are insolidsolutionandthe eects
from the dierent solute atoms are additive, solid-solution
strengthening in this alloy accounts for a strength increase
of 82 MPa. Because the actual contribution from solid-
solution strengthening is <82 MPa in the T6-tempered sam-
ples, where a signicant fraction of the solute atoms have
precipitated, we conclude that solid-solution strengthening
provides a small contribution to the total strength of the
as-extruded and T6-tempered materials.
4.3. Dislocation strengthening
Dislocations interact with themselves and impede their
own motion. Thus, increasing the dislocation density in a
Table 7
Estimated strength increment for dierent strengthening mechanisms.
UFG7075-
E
UFG7075-E-
T6
CG7075-
E
CG7075-E-
T6
Dr
grain boundary
(MPa)
242 185 127 120
Dr
solid solution
(MPa)
<82 82 <82 82
Dr
dislocation
(MPa) 99 95 61 35
Dr
orowan
(MPa) 45 414 102 472
Table 8
Physical meaning and values of dierent symbols used in the strengthening
mechanism calculations [25,36,43,4650].
Symbol Meaning Values Unit
a Lattice constant =0.405 for fcc Al nm
b Magnitude of the
Burgers vector
=
p
2/2a = 0.286 for fcc
metals
nm
k
y
HallPetch
coecient
=0.12 MPa=

m
p
M Mean orientation
factor
=3.06 for the fcc
polycrystalline matrix
Dimensionless
G Shear modulus =26.9 for Al 7075 GPa
a Constant =0.2 for fcc metals Dimensionless
t Poisson ratio =0.33 for Al 7075 Dimensionless
a
e
Constant =2.6 for fcc metals Dimensionless
K. Ma et al. / Acta Materialia xxx (2013) xxxxxx 11
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
metal increases the yield strength of the material [47]. To
evaluate and compare the role of the residual dislocations
to strengthening in the CG and UFG 7075 materials, the
BaileyHirsch relationship was applied in the current study
[47,64]:
Dr
d
MaGbq
1
2
: 6
The meaning and values of the symbols in Eq. (6) are
given in Table 8. The applicability of the BaileyHirsch
relationship to UFG materials is supported by previous
work by other researchers [25,35,45,65]. However, it is
noteworthy that Huang et al. [66] reported an interesting
phenomenon: hardening by annealing and softening by
deforming for a nanostructured Al prepared by ARB,
which is in contrast to the typical behavior of a metal. They
proposed that many dislocation sinks available in the form
of closely spaced high-angle boundaries reduce the number
of dislocation sources during annealing. Consequently, an
increase in the yield stress is expected during straining in
order to activate new dislocation sources. In the present
work, however, we aimed to estimate the contribution from
the residual dislocations to strengthening during tensile
deformation. The dislocation density values were deter-
mined by XRD (Section 3.2.3). The strength increment
caused by dislocation sources is calculated to be 99 and
95 MPa in samples UFG7075-E and UFG7075-E-T6,
respectively. In contrast, dislocation strengthening contrib-
uted an increase of 61 and 35 MPa for samples
CG7075-E and CG7075-E-T6, respectively. An additional
strengthening mechanism related to dislocations, the so-
called dislocation source-limited hardening, may also oper-
ate for UFG, but not CG, materials, as a higher stress is
required to activate alternative dislocation sources
[56,66]. Combined, these results suggest that dislocation
strengthening plays a more signicant role in the UFG
7075 materials than in the CG 7075 materials.
4.4. Precipitate and dispersoid strengthening
Precipitates (GP zones, g
0
- and g-phases) are present in
both CG and UFG 7075 materials, in both the as-extruded
and the T6-tempered conditions. Nitrogen- or oxygen-rich
dispersoids are only present in the UFG 7075 materials
because they are a by-product of cryomilling [14,15,22].
Precipitation and dispersoid strengthening are governed
by either the Orowan dislocation bypassing or dislocation
shearing mechanisms. The one causing a smaller strength
increment is the operative mechanism [25,47]. When pre-
cipitates or dispersoids are bypassed by the Orowan dislo-
cation bypassing mechanism, the yield strength increment,
Dr
orowan
, is [46,47,67]:
Dr
orowan
M
0:4Gb
p

1 t
p
ln2r=b
k
p
; 7
where M, G, b, t are dened in Table 8; r is the mean radius
of a circular cross-section in a random plane for a spherical
precipitate, r

2=3
p
r, where r is the mean radius of the
precipitates. In the shearing mechanism, three factors con-
tribute to the increase in yield strength: coherency strength-
ening (Dr
cs
), modulus mismatch strengthening (Dr
ms
) and
order strengthening (Dr
os
). The larger of Dr
cs
+ Dr
ms
or
Dr
os
is the total strength increment from the dislocation
shearing mechanism [25,46,47]. The values of Dr
cs
, Dr
ms
and Dr
os
are calculated from Eqs. (8)(10), respectively
[25,43,68,69]:
Dr
cs
Ma
e
Ge
c

3
2
rf
0:5Gb
1
2
; 8
Dr
ms
M0:0055DG
3
2
2f
G
1
2
r
b
3m
2
1
; 9
Dr
os
M0:81
c
apb
2b
3pf
8
1
2
; 10
where M, G, b and r are listed in Table 8; a
e
= 2.6 for face-
centered cubic (fcc) metals; m = 0.85; DG is the modulus
mismatch between the matrix and the precipitates; e
c
is
the constrained lattice parameter mist; f is the volume
fraction of the precipitates; and c
apb
is the antiphase
boundary free energy of the precipitate phase.
To estimate the strength increment from the precipitates
and dispersoids in the CG and UFG 7075 materials, the
operative mechanism for each type of precipitate and dis-
persoid must rst be identied. For precipitates, e.g. the
g-phase precipitates that are incoherent with the Al-
matrix, the operative mechanism is Orowan dislocation
bypassing [1,6,70]. If the precipitate is coherent or semico-
herent with the matrix, e.g. GP zones and the g
0
-phase pre-
cipitates, the strength increment resulting from dislocation
shearing needs to be evaluated and compared with Dr
orowan
to determine the operative mechanism [25,47]. From Eq.
(7), it is evident that Dr
orowan
is only dependent on r and
k
p
, and is independent of the intrinsic properties of the pre-
cipitates or dispersoids, e.g. chemical composition and
crystal structure [47]. The strength increment from disloca-
tion shearing (Eqs. (8)(10)) is, however, dependent on the
intrinsic material properties, which derive from chemical
composition and crystal structure [46,47,67]. Due to the
challenges of identifying the chemical composition and
crystal structure of each precipitate and dispersoid from
TEM images, the upper bound of Dr
orowan
was estimated
Table 9
Data on the primary solute atoms in Al 7075 and their contribution to
yield strength.
Element Dierence in
atomic
radii [50,63],
(r
x
r
Al
)/r
Al
(%)
Yield strength
addition
[50,63]
(MPa wt.%
1
)
Concentration
(wt.%)
Contribution
to yield
strength
(MPa)
Zn 6 2.9 5.4 16
Mg 11.8 18.6 2.4 44
Cu 10.7 13.8 1.6 22
12 K. Ma et al. / Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
by assuming that Orowan dislocation bypassing is opera-
tive for all the precipitates and dispersoids (GP zones,
g-phase precipitates, g
0
-phase precipitates and MgO dis-
persoids). This is only the starting point since the actual
value of Dr
orowan
for individual precipitate/dispersoid types
will be lower than this upper bound, and it is possible that
Orowan dislocation bypassing is not the operative
mechanism for each precipitate or dispersoid type. Using
the values of r and k
p
, Table 5, in Eq. (7), the upper bound
for the Dr
orowan
for the aggregate eect of all the precipi-
tates and dispersoids, as well as for each precipitate and
dispersoid type, in samples UFG7075-E, UFG7075-E-T6,
CG7075-E and CG7075-E-T6, are estimated to be 45,
414, 102 and 472 MPa, respectively.
Next, for comparison and identication of the operable
strengthening mechanisms, the strength increments from
dislocation shearing for g
0
-phase precipitates, GP zones
and MgO dispersoids are estimated. This mechanism is
not a viable option for the g-phase precipitates because
they are incoherent with the Al matrix. The relationship
between f, k
p
and r is [71]:
k
p
2r

p
4f
r
1

: 11
Substituting f for k
p
using Eq. (11), Eq. (8) becomes:
Dr
cs

3
2
1
4

2p
b
r
MGa
e
e
c
r
3
2
k
p
2r
: 12
The g
0
-MgZn
2
phase has an hexagonal structure:
a = 0.496 nm and c = 1.402 nm [72]. The orientation rela-
tionship between the g
0
-phase and the Al matrix is
(001)
g
0 //{111}
Al
and [110]
g
0 //h112i
Al
[72]. The interrela-
tionship between the lattice parameters of the g
0
-phase
and the Al matrix are given by d
100
(g
0
) = 3d
220
(Al) and
d
001
(g
0
) = 6d
111
(Al), which makes the g
0
-semicoherent with
the Al fcc lattice [72]. To calculate the semicoherent lattice
parameter mist between precipitates (hexagonal structure)
and matrix (cubic structure), the unique constrained eec-
tive mist strain, e
c
, is dened by [25,73]:
e
c

1 t
31 t
e
eff
1 4G=3B
c

; 13
where B
c
is the bulk modulus of the g
0
-MgZn
2
precipitates,
63.5 GPa [74], and e
e
is:
e
eff

2
p
3
e
11
e
22

2
e
22
e
33

2
e
33
e
11

1
2
: 14
where
e
11

a
p
a
m
1; e
22

3
p
a
p

2
p
a
m
1; e
33

c
p

2
p
a
m
1; 15
where a
m
is the lattice parameter of the Al matrix,
0.405 nm; a
p
and c
p
are the lattice parameters of the
g
0
-phase precipitate: a
p
= 0.496 nm and c
p
= 1.402 nm,
respectively [72]. Utilizing Eqs. (11)(13), the value of e
c
is 0.3. Substituting e
c
and the values of r for the g
0
-phase
and the values of k
p
into Eq. (12), the lower bound of the
strength increments due to coherency strengthening,
Dr
cs
, are approximately several GPa for all the materials.
Thus, the upper bound of the strength increment from Oro-
wan dislocation bypassing is smaller than the lower bound
of strengthening from coherency for all cases, i.e.
Dr
orowan
< Dr
cs
, implying that Orowan dislocation bypass-
ing is the operative mechanism for the g
0
-phase because the
smaller of Dr
orowan
and Dr
shearing
is the operative mecha-
nism [25,47].
In the case of GP zones, Berg
0
s study [8] revealed that
GP zones are Zn-rich layers on {111} planes, with internal
order in the form of elongated h110i domains and a spac-
ing between rows of atoms 68% less than in the Al
matrix. The reduced spacing, relative to the Al matrix lay-
ers above or below the sheet, was proposed to be associated
with the smaller radius of Zn atoms [8]. In Eq. (12), we
assumed the e
c
came from this reduced spacing and is equal
to 6%. Utilizing the values of r for GP zones and k
p
into
Eq. (12), the lower bound of Dr
cs
for GP zones in samples
UFG7075-E, UFG7075-E-T6 and CG7075-E-T6 are
approximately 96, 1000 and 590 MPa, respectively. These
values are greater than the upper bound of Dr
orowan
, i.e.
Dr
orowan
< Dr
cs
. Hence, similar to the g
0
-phase, Orowan
dislocation bypassing is determined to be the operative
mechanism for GP zones.
For the UFG 7075 materials, oxide dispersoids must
also be considered because cryomilling promotes their for-
mation [14,15]. The APT analyses demonstrate that the
volume fraction of MgO dispersoids is 0.14% in
UFG7075-E, with an average diameter of 4 nm. Using
Eq. (11), k
MgO
is 73 nm, the bulk modulus of MgO is
155 GPa [75] and MgO has a cubic structure with a lattice
parameter of 4.21 A

[76]. The lattice parameter mist


between MgO and Al, e, is 0.04. Applying Eq. (12), the
coherency strengthening, Dr
cs
, from MgO, is 220 MPa.
This value is larger than the upper bound of Dr
orowan
in
UFG7075-E (45 MPa). Therefore, Orowan dislocation
bypassing is identied as the operative strengthening mech-
anism for MgO dispersoids. Although sample UFG7075-
E-T6 was not characterized by APT, it is reasonable to
assume that strengthening from oxide dispersoids is similar
to that in sample UFG7075-E, because the thermal treat-
ments from solutionizing and T6 temper are not expected
to modify the oxide dispersoid distribution characteristics.
Consequently, Orowan dislocation bypassing has been
determined to be the operative strengthening mechanism
for all of the second-phase particles (the three families of
precipitates and the oxide dispersoids) in all the materials,
which leads to strength increments of 45, 414, 102 and
472 MPa for samples UFG7075-E, UFG7075-E-T6,
CG7075-E and CG7075-E-T6, respectively. Comparing
these calculated values, it is clear that precipitation
strengthening plays a more signicant role in the CG
7075 materials than in UFG 7075 for equivalent condi-
tions. The experimental results, Table 4, suggest that sam-
ple UFG7075-E-sol maintained the same strength as that
K. Ma et al. / Acta Materialia xxx (2013) xxxxxx 13
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
of sample UFG7075-E. We speculate that the loss of pre-
cipitation strengthening (45 MPa) after solution treatment
was compensated for by the increase from solid-solution
strengthening because 45 MPa is within the range of the
ideal solid-solution strengthening increment with an upper
limit of 82 MPa. In contrast, sample CG7075-E-sol
exhibited a strength decrease of 47 MPa, compared with
the strength of sample CG7075-E. Comparing the calcu-
lated values, the loss of precipitation strengthening
(102 MPa) in sample CG7075-E after solution treatment
cannot be totally compensated for by solid-solution
strengthening, less than 82 MPa, which is consistent with
our experimental observations.
5. Conclusions
The mechanical properties of UFG Al 7075 materials,
fabricated by cryomilling feedstock powder with subse-
quent consolidation and various heat-treatment conditions,
were investigated, with direct comparison to those of CG
7075 counterpart materials that were consolidated and heat
treated through identical processes using commercially gas-
atomized powder as the feedstock. The microstructures of
samples in the as-extruded and the T6 temper condition
were characterized by XRD and TEM. Additionally,
APT was utilized to study the UFG 7075 material to inves-
tigate the nanoscale precipitates and oxide dispersoids that
are produced during cryomilling. The contributions from
dierent strengthening mechanisms were quantitatively
measured (including grain-boundary (HallPetch strength-
ening), dislocation, solid-solution, precipitation and oxide
dispersoid strengthening) in the UFG 7075 material in
the as-extruded condition and after the T6 treatment, in
comparison with the CG 7075 counterparts. With multiple
families of precipitates and oxide dispersoids forming in the
materials, Orowan dislocation bypassing is the dominant
strengthening mechanism in the CG 7075 materials, espe-
cially after the T6 temper. Grain-boundary strengthening
is dominant in the as-extruded UFG 7075, and it also
makes a signicant contribution in sample UFG7075-E-
T6. In summary, we conclude the following:
(i) The UFG 7075 materials exhibit higher strength than
the CG 7075 materials in each equivalent condition.
The strength of the as-extruded UFG 7075 (YS:
583 MPa, UTS: 631 MPa) is even higher than com-
mercial Al 7075 with the T6 temper (YS: 541 MPa,
UTS: 593 MPa) and is close in value to the CG
7075 sample with the T6 temper (YS: 613 MPa,
UTS: 659 MPa).
(ii) After a T6 temper, both the UFG and CG 7075 mate-
rials exhibit a considerable increase in strength. The
YS and UTS of sample UFG7075-E-T6 are 734 and
774 MPa, respectively, which are 120 MPa higher
than those of sample CG7075E-T6 (YS: 613 MPa,
UTS: 659 MPa). The age-hardening eect, i.e. the
strength increment after natural aging or T6 temper,
in the UFG 7075 materials is not, however, as signif-
icant as that in the CG 7075 materials. The increases
in yield strength of samples CG7075-E-Nag and
CG7075-E-T6 are 70% and 160%, respectively, rela-
tive to that of CG7075-E-sol. In contrast, the
increases in yield strength after natural aging and
T6 temper for the UFG 7075 materials are only
13% and 26%, respectively.
(iii) Sample UFG7075-E possesses a relatively small vol-
ume fraction, 0.07% from APT, of GP zones, and
g
0
-phase precipitates in grain interiors: the average
diameter of both precipitate types is <5 nm. APT
results demonstrate that this material also contains
0.14% volume fraction of MgO dispersoids, with an
average diameter of 4 nm. In contrast, sample
CG7075-E exhibits a number of plate-like g
0
-phase
precipitates (55 nm diameter) and lath-like g-phase
precipitates (length 174 nm) in grain interiors: these
precipitates pin dislocations. After the T6 temper, the
grains in the UFG 7075 material contain a volume
fraction 1.5% of spherical GP zones and platelet-
like g
0
-phase precipitates, both in nanoscale (diame-
ter 6 5 nm), while sample CG7075-E-T6 exhibits
many spherical GP zones, diameter < 5 nm, but coar-
ser g
0
-phase precipitates whose average diameter is
61 nm.
(iv) The analyses of the contributions from dierent
strengthening mechanisms indicate that grain-bound-
ary strengthening is the predominant mechanism in
the as-extruded UFG 7075 material, contributing a
strength increment of 242 MPa, and it is also an
important contribution (185 MPa) in the T6-tem-
peredUFG7075; while precipitationOrowanstrength-
ening is predominant in the T6-tempered UFG 7075,
contributing a strength increment of 414 MPa. In con-
trast, Orowan strengthening is the primary strength
contributor in the CG 7075 materials, both in the as-
extruded and T6-tempered states, with values of
approximately 102 and 472 MPa, respectively.
Acknowledgements
The authors would like to acknowledge nancial sup-
port provided by the Oce of Naval Research (Grant
No. ONR N00014-12-1-0237), Dr. Lawrence Kabaco as
the Program Ocer. The authors are also grateful for tech-
nical discussions with Dr. Ali Youseani from Boeing Re-
search & Technology. Atom-probe tomography was
performed at the Northwestern University Center for
Atom-Probe Tomography (NUCAPT). The local-electrode
atom-probe (LEAP) tomograph was purchased and up-
graded with funding from NSF-MRI (DMR-0420532)
and ONR-DURIP (N00014-0400798, N00014-0610539,
N00014-0910781) grants. Instrumentation at NUCAPT
was also supported by the Initiative for Sustainability
and Energy at Northwestern (ISEN). This research also
14 K. Ma et al. / Acta Materialia xxx (2013) xxxxxx
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042
made use of the Shared Facilities at the Materials Research
Center of Northwestern University, supported by the
National Science Foundations MRSEC program
(DMR-1121262).
References
[1] Fine M. Metall Mater Trans A 1975;6:625.
[2] Nakai M, Eto T. High strength heat treatable 7000 series aluminum
alloy of excellent corrosion resistance and a method of producing
thereof. United States: Kabushiki Kaisha Kobe Seiko Sho (Kobe,
JP), US Patent 6048415, 2000.
[3] Park JK, Ardell AJ. Mater Sci Eng A 1989;114:197.
[4] Zhao YH, Liao XZ, Jin Z, Valiev RZ, Zhu YT. Acta Mater
2004;52:4589.
[5] Viana F, Pinto AMP, Santos HMC, Lopes AB. J Mater Process
Technol 1999;9293:54.
[6] Richard D, Adler P. Metall Mater Trans A 1977;8:1177.
[7] Adler P, DeIasi R. Metall Mater Trans A 1977;8:1185.
[8] Berg LK, Gjonnes J, Hansen V, Li XZ, Knutson-Wedel M, Waterloo
G, et al. Acta Mater 2001;49:3443.
[9] Zhao YH, Liao XZ, Zhu YT, Valiev aRZ. J Mater Res 2005;20:288.
[10] Chen H, Yang B. Mater Trans 2008;49:2912.
[11] Donoso E. Mater Sci Eng 1985;74:39.
[12] Gleiter H. Nanostruct Mater 1995;6:3.
[13] Gleiter H. Acta Mater 2000;48:1.
[14] Lavernia EJ, Han BQ, Schoenung JM. Mater Sci Eng A
2008;493:207.
[15] Witkin DB, Lavernia EJ. Prog Mater Sci 2006;51:1.
[16] Zhu Y, Langdon T. JOM J Miner Metal Mater Soc 2004;56:58.
[17] Zhao YH, Liao XZ, Cheng S, Ma E, Zhu YT. Adv Mater
2006;18:2280.
[18] Panigrahi SK, Jayaganthan R. Mater Des 2011;32:3150.
[19] Panigrahi S, Jayaganthan R. Metall Mater Trans A 2011;42:3208.
[20] Liddicoat PV, Liao X-Z, Zhao Y, Zhu Y, Murashkin MY, Lavernia
EJ, et al. Nat Commun 2010;1:63.
[21] Pickens JR, Christodoulou L. Metall Trans A 1987;18:135.
[22] Ma K, Schoenung JM. Philos Mag Lett 2010;90:739.
[23] Tang F, Hagiwara M, Schoenung JM. Mater Sci Eng A 2005;407:306.
[24] Zhang Z, Topping T, Li Y, Vogt R, Zhou Y, Haines C, et al. Scripta
Mater 2011;65:652.
[25] Wen H, Topping TD, Isheim D, Seidman DN, Lavernia EJ. Acta
Mater 2013;61:2769.
[26] Li Y, Zhang Z, Vogt R, Schoenung JM, Lavernia EJ. Acta Mater
2011;59:7206.
[27] Topping T, Ahn B, Li Y, Nutt S, Lavernia E. Metall Mater Trans A
2012;43:505.
[28] Miller MK, Cerezo A, Hetherington MG, Smith GDW. Atom-probe
eld-ion microscopy. Oxford: Clarendon Press; 1996.
[29] Seidman DN. Annu Rev Mater Res 2007;37:127.
[30] Seidman DN, Stiller K. MRS Bull 2009;34:717.
[31] Kelly TF, Miller MK. Rev Sci Instrum 2007;78:031101.
[32] Blavette D, Deconihout B, Bostel A, Sarrau JM, Bouet M, Menand
A. Rev Sci Instrum 1993;64:2911.
[33] Mulholland MD, Seidman DN. Acta Mater 2011;59:1881.
[34] Ringer SP, Hono K. Mater Charact 2000;44:101.
[35] Wang JS, Mulholland MD, Olson GB, Seidman DN. Acta Mater
2013;61:4939.
[36] Metals handbook, properties and selection: nonferrous alloys and
special-purpose materials, 10th ed. Materials Park, OH: ASM
International; 1990.
[37] Zhang Z, Dallek S, Vogt R, Li Y, Topping TD, Zhou Y, et al. Metall
Mater Trans A 2010;41:532.
[38] Hu T, Ma K, Topping TD, Schoenung JM, Lavernia EJ. Acta Mater
2013;61:2163.
[39] Williamson GK, Hall WH. Acta Metall 1953;1:22.
[40] Collins TJ. Biotechniques 2007;43:S25.
[41] Hellman OC, Vandenbroucke JA, Ru sing J, Isheim D, Seidman DN.
Microsc Microanal 2000;6:437.
[42] Topping TD, Lavernia EJ. Strain hardening, strain softening and the
Portevin-Le Chatelier eect in cryomilled, ultrane grained AA 5083.
In: Weiland H, Rollett AD, Cassada WA, editors. ICAA13: 13th
international conference on aluminum alloys. Hoboken, NJ: Cass;
2012.
[43] Seidman DN, Marquis EA, Dunand DC. Acta Mater 2002;50:4021.
[44] Li Y, Liu W, Ortalan V, Li WF, Zhang Z, Vogt R, et al. Acta Mater
2010;58:1732.
[45] Cao B, Joshi SP, Ramesh KT. Scripta Mater 2009;60:619.
[46] Kim YW, Grith WM. In: Dispersion strengthened aluminum alloys:
proceedings of the six session symposium on dispersion strengthened
aluminum alloys. Phoenix, AZ: TMS; 1988. p. 157.
[47] Courtney TH. Mechanical behavior of materials. Long Grove: Wave-
land; 2005.
[48] Kim YW, Grith WM. PM aerospace materials. Shrewsbury: MPR;
1984.
[49] Wert JA. Strength of metal alloys. In: Gifkins RC, editor. Oxford:
Pergamon Press; 1980. p. 339.
[50] J.R. Davis & Associates, ASM international handbook committee.
Aluminum and aluminum alloys. ASM specialty handbook. Materi-
als Park, OH: ASM, International; 1993. p. 784.
[51] Malygin GA. Phys Usp 2011;54:1091.
[52] Pande CS, Cooper KP. Prog Mater Sci 2009;54:689.
[53] Hall EO. Proc Phys Soc Sec B 1951;64:747.
[54] Petch NJ. J Iron Steel Inst Lond 1953;174:25.
[55] Loucif A, Figueiredo RB, Baudin T, Brisset F, Chemam R, Langdon
TG. Mater Sci Eng A 2012;532:139.
[56] Kamikawa N, Huang X, Tsuji N, Hansen N. Acta Mater
2009;57:4198.
[57] Furukawa M, Horita Z, Nemoto M, Valiev RZ, Langdon TG. Acta
Mater 1996;44:4619.
[58] Hayes RW, Witkin D, Zhou F, Lavernia EJ. Acta Mater
2004;52:4259.
[59] Carsley JE, Fisher A, Milligan WW, Aifantis EC. Metall Mater Trans
A 1998;29:2261.
[60] Fleischer RL. Acta Metall 1962;10:835.
[61] Fleischer RL. Acta Metall 1963;11:203.
[62] Rupert TJ, Trenkle JC, Schuh CA. Acta Mater 2011;59:1619.
[63] Topping TD. PhD dissertation: Nanostructured Aluminum Alloys
and Their Composites, University of California Davis; 2012.
[64] Bailey JE, Hirsch PB. Philos Mag 1960;5:485.
[65] Kumar N, Mishra RS. Mater Sci Eng A 2013;580:175.
[66] Huang X, Hansen N, Tsuji N. Science 2006;312:249.
[67] Srivatsan TS. Mater Manuf Process 1991;6:565.
[68] Ardell A. Metall Mater Trans A 1985;16:2131.
[69] Booth-Morrison C, Dunand DC, Seidman DN. Acta Mater
2011;59:7029.
[70] Fribourg G, Brechet Y, Deschamps A, Simar A. Acta Mater
2011;59:3621.
[71] Brown L, Ham R. Strengthening methods in crystals. Amster-
dam: Elsevier; 1971. p. 9.
[72] Li XZ, Hansen V, Gjonnes J, Wallenberg LR. Acta Mater
1999;47:2651.
[73] Mulholland MD. PhD thesis, Materials Science and Engineering,
Northwestern University; 2011.
[74] Wu M-M, Wen L, Tang B-Y, Peng L-M, Ding W-J. J Alloy Compd
2010;506:412.
[75] Madelung O, Rossler U, Schulz M, editors. Magnesium oxide (MgO)
Youngs, shear and bulk moduli, Poissons ratio, vols. III/17B-22A-
41B. Springer MaterialsThe Landolt-Bornstein, Database. doi:
http://dx.doi.org/10.1007/1081719_210.
[76] Madelung O, Ro ssler U, Schulz M, editors. Magnesium oxide (MgO)
crystal structure, lattice parameters, thermal expansion, vols. III/17B-
22A-41B. SpringerMaterialsThe Landolt-Bo rnstein, Database. doi:
http://dx.doi.org/10.1007/10681719_206.
K. Ma et al. / Acta Materialia xxx (2013) xxxxxx 15
Please cite this article in press as: Ma K et al. Mechanical behavior and strengthening mechanisms in ultrane grain precipitation-
strengthened aluminum alloy. Acta Mater (2013), http://dx.doi.org/10.1016/j.actamat.2013.09.042

You might also like